Content uploaded by John Geissman
Author content
All content in this area was uploaded by John Geissman on Mar 14, 2016
Content may be subject to copyright.
The Geological Society of America Geologic Time Scale
J.D. Walker1,†, J.W. Geissman2, S.A. Bowring3, and L.E. Babcock4
1Department of Geology, University of Kansas, Lawrence, Kansas 66045, USA
2Department of Geosciences, ROC 21, University of Texas at Dallas, Richardson, Texas 75080, USA, and
Department of Earth and Planetary Sciences, MSC 03 2040, 1 University of New Mexico, Albuquerque, New Mexico 87131, USA
3Department of Earth, Atmospheric, and Planetary Sciences, Massachusetts Institute of Technology, Cambridge,
Massachusetts 02139, USA
4Department of Geology, Lund University, SE-223 62 Lund, Sweden, and School of Earth Sciences, Ohio State University,
Columbus, Ohio 43210, USA
ABSTRACT
The Geological Society of America has
sponsored versions of the geologic time scale
since 1983. Over the past 30 years, the Geo-
logical Society of America Geologic Time
Scale has undergone substantial modifi ca-
tions, commensurate with major advances
in our understanding of chronostratig-
raphy, geochronology, astrochronology,
chemostratigraphy, and the geomagnetic
polarity time scale. Today, many parts of
the time scale can be calibrated with preci-
sions approaching less than 0.05%. Some
notable time intervals for which collabora-
tive, multifaceted efforts have led to dra-
matic improvements in our understanding
of the character and temporal resolution
of key evolutionary events include the Tri-
assic-Jurassic, Permian-Triassic, and Neo-
proterozoic-Phanerozoic boundaries (or
transitions). In developing the current Geo-
logical Society of America Time Scale, we
have strived to maintain a consistency with
efforts by the International Commission on
Stratigraphy to develop an international
geologic time scale.
Although current geologic time scales are
vastly improved over the fi rst geologic time
scale, published by Arthur Holmes in 1913,
we note that Holmes, using eight numeri-
cal ages to calibrate the Phanerozoic time
scale, estimated the beginning of the Cam-
brian Period to within a few percent of the
currently accepted value. Over the past 100
years, the confl uence of process-based geo-
logical thought with observed and approxi-
mated geologic rates has led to coherent and
quantitatively robust estimates of geologic
time scales, reducing many uncertainties to
the 0.1% level.
INTRODUCTION
One of the most important aspects of research
in the geosciences is connecting what we examine
in the rock record with ages of events and mea-
sured rates and durations of geologic processes .
In doing so, geoscientists are able to place esti-
mates on the rates of climate and evolutionary
changes, use astronomically forced depositional
processes to tell time within sedimentary basins,
examine the ways in which tectonic processes
change crustal and mantle structure and infl uence
landscape evolution and global climate patterns,
and assess the temporal relations among magma-
tism, fl uid-rock interaction, and base/precious
metal mineralization in many settings. A key re-
quirement is establishing accurate ages of rocks
that are directly associated with or bracket a geo-
logic event or process. With efforts to estimate
the chronology of geologic events beginning well
over two centuries ago, this was commonly ac-
complished by establishing a relative geologic
time scale using the ranges of fossils and strati-
graphic relationships. Beginning in the early
twentieth century, the relative geologic time scale
was calibrated using numerical information.
A geologic time scale is the ordered com-
pilation of numerical ages and relative age de-
terminations based on stratigraphic and other
principles. Numerical ages, formerly called
“absolute ages” (Holmes, 1962), form a chrono-
metric time scale typically expressed in thou-
sands (ka) or millions (Ma) of years; relative
ages form a chronostratigraphic time scale. A
geologic time scale is an invaluable tool for geo-
scientists investigating virtually any aspect of
Earth’s development, anywhere on the planet,
and at almost any time in Earth’s history.
This paper describes the history of the devel-
opment of the Geological Society of America
Geologic Time Scale and provides a brief his-
tory of geologic time scales and their compo-
nents. We also discuss important advances made
over the past few decades in establishing both
numerical and relative ages, describe selected
proxies for time in the rock record, and note and
comment on some future challenges. This paper
is intended to provide a general overview of
geologic time scales. The most comprehensive
treatment of the geologic time scale is contained
in the recent publication of Gradstein et al.
(2012), the most current defi nitive work on the
geologic time scale from a global perspective.
This book is the most recent in the series of ma-
jor publications by The Geological Society of
London (Harland et al., 1964) and subsequently
Cambridge University Press (Harland et al.,
1982, 1990; Gradstein et al., 2004; Ogg et al.,
2008) and Elsevier (Gradstein et al., 2012). The
current Geological Society of America Geologic
Time Scale (Fig. 1) incorporates information
presented in the International Commission on
Stratigraphy’s International Chronostratigraphic
Chart (Cohen et al., 2012) and in Gradstein
et al. (2012). Numerical dates used for bound-
ary positions are from Gradstein et al. (2012).
The Geological Society of America (GSA) does
not directly “maintain” an international geo-
logic time scale. Rather, the society provides a
geologic time scale in a concise, logically orga-
nized and readable format that is largely based
on the work of the International Commission on
Stratigraphy (ICS) and related groups and pub-
lications. GSA follows the work and recommen-
dations of these groups in promoting a better
understanding and use of geologic time through
its time scale. Many of these organizations in-
clude geoscientists who are GSA members or
members of the associated societies of GSA.
History of Chronometric-
Chronostratigraphic Geologic Time Scales
“For it was evident to me that the space between
the mountain ranges, which lie above the City of
Memphis, once was a gulf of the sea, like the regions
For permission to copy, contact editing@geosociety.org
© 2013 Geological Society of America
259
GSA Bulletin; March/April 2013; v. 125; no. 3/4; p. 259–272; doi:10.1130/B30712.1; 1 fi gure; 1 table.
†E-mail: jdwalker@ku.edu
Invited Review
CELEBRATING ADVANCES IN GEOSCI ENCE
1888 2013
on August 30, 2015gsabulletin.gsapubs.orgDownloaded from
Walker et al.
260 Geological Society of America Bulletin, March/April 2013
HIST
ANOM.
CHRON.
C31
C32
C33
31
32
33
M0r
M1
M5
M10
M12
M14
M16
M18
M20
M22
M25
M29
M3
RAPID POLARITY CHANGES
30 C30
C34
34
1C1
C2
C2A
C3
C3A
C4
C4A
C6
C6A
C6B
C6C
C7
C8
C9
C10
C11
C12
C13
C15
C16
C17
C18
C19
C20
C21
C22
C23
C24
C25
C26
C27
C28
C29
C7A
C5
C5A
C5B
C5C
C5D
C5E
2
2A
3
3A
4
4A
5
5B
5A
5C
6
6A
6B
7
8
9
10
11
12
13
15
16
17
18
19
20
21
22
23
24
25
28
29
26
27
7A
6C
5D
5E
30 C30
237
70
80
90
100
110
120
130
140
150
160
170
180
190
210
200
220
230
240
250
5
10
15
20
25
30
35
40
45
50
55
60
65
750
1000
1250
1500
1750
2000
2250
2500
2750
3000
3250
3500
3750
260
280
300
320
340
380
360
400
420
440
460
480
500
520
540
4000
GEOLOGIC TIME SCALE
PALEOZOIC
PERMIAN
DEVONIAN
ORDOVICIAN
SILURIAN
MISSIS-
SIPPIAN
PENNSYL-
VANIAN
CAMBRIAN CARBONIFEROUS
AGE
(Ma) EPOCH AGE PICKS
(Ma)
PERIOD
252
260
254
265
269
272
279
290
296
304
307
299
323
331
347
359
372
383
388
393
408
411
419
423
426
433
430
439
441
427
444
445
453
458
470
467
478
485
494
497
501
505
490
509
514
521
529
541
GZHELIAN
KASIMOVIAN
MOSCOVIAN
BASHKIRIAN
SERPUKHOVIAN
VISEAN
TOURNAISIAN
FAMENNIAN
FRASNIAN
GIVETIAN
EIFELIAN
EMSIAN
PRAGIAN
LOCHKOVIAN
315
PRECAMBRIAN
PROTEROZOICARCHEAN
AGE
(Ma) EON ERA
BDY.
AGES
(Ma)
1000
1200
1800
2050
2300
1400
1600
2500
2800
3200
3600
4000
Lopin-
gian
MIDDLE
Guada-
lupian
Cisura-
lian
LLANDO-
VERY
EARLY
EARLY
FURON-
GIAN
Epoch 3
Epoch 2
TERRE-
NEUVIAN
LATE
LUDLOW
LATE
MIDDLE
WENLOCK
541
635
850
MESOZOIC
TRIASSIC JURASSIC CRETACEOUS
AGE
(Ma) EPOCH AGE PICKS
(Ma)
MAGNETIC
POLARITY PERIOD
PERIOD
LATE
EARLY
LATE
EARLY
MIDDLE
LATE
EARLY
MIDDLE
MAASTRICHTIAN
66.0
72.1
83.6
86.3
89.8
93.9
100
113
126
131
134
139
145
152
157
166
164
CAMPANIAN
SANTONIAN
CONIACIAN
TURONIAN
CENOMANIAN
ALBIAN
APTIAN
BARREMIAN
HAUTERIVIAN
VALANGINIAN
BERRIASIAN
TITHONIAN
KIMMERIDGIAN
OXFORDIAN
CALLOVIAN
BATHONIAN
BAJOCIAN
AALENIAN
TOARCIAN
PLIENSBACHIAN
SINEMURIAN
HETTANGIAN
NORIAN
RHAETIAN
CARNIAN
LADINIAN
ANISIAN
OLENEKIAN
INDUAN
168
170
174
199
191
201
209
241
228
252
250
247
RAPID POLARITY CHANGES
CENOZOIC
AGE
(Ma) EPOCH AGE PICKS
(Ma)
MAGNETIC
POLARITY PERIOD
HIST.
ANOM.
CHRON.
QUATER-
NARY PLEISTOCENE*
MIOCENEOLIGOCENEEOCENEPALEOCENE
PLIOCENE PIACENZIAN
0.01
1.8
3.6
5.3
7.2
11.6
13.8
16.0
20.4
23.0
28.1
33.9
37.8
41.2
47.8
56.0
59.2
61.6
66.0
ZANCLEAN
MESSINIAN
TORTONIAN
SERRAVALLIAN
LANGHIAN
BURDIGALIAN
AQUITANIAN
CHATTIAN
RUPELIAN
PRIABONIAN
BARTONIAN
LUTETIAN
YPRESIAN
DANIAN
THANETIAN
SELANDIAN
CALABRIAN
HOLOCENE
PALEOGENE NEOGENE
GELASIAN 2.6
183
CHANGHSINGIAN
WORDIAN
ROADIAN
WUCHIAPINGIAN
CAPITANIAN
KUNGURIAN
ASSELIAN
SAKMARIAN
ARTINSKIAN
PRIDOLI
LUDFORDIAN
GORSTIAN
HOMERIAN
RHUDDANIAN
TELYCHIAN
AERONIAN
SHEINWOODIAN
HIRNANTIAN
SANDBIAN
KATIAN
DARRIWILIAN
DAPINGIAN
AGE 10
JIANGSHANIAN
PAIBIAN
GUZHANGIAN
DRUMIAN
AGE 5
AGE 4
AGE 3
AGE 2
FORTUNIAN
FLOIAN
TREMADOCIAN
EDIACARAN
CRYOGENIAN
TONIAN
STENIAN
ECTASIAN
CALYMMIAN
STATHERIAN
OROSIRIAN
RHYACIAN
SIDERIAN
NEOPRO-
TEROZOIC
MESOPRO-
TEROZOIC
PALEOPRO-
TEROZOIC
NEOARCHEAN
MESO-
ARCHEAN
PALEO-
ARCHEAN
EOARCHEAN
HADEAN
EARLY
EARLY
MIDDLE
MIDDLE
LATE
LATE
Figure 1. Geological Society of America Geologic Time Scale. The Cenozoic, Mesozoic, and Paleozoic are the Eras of the Phanerozoic Eon. Names of
chrono stratigraphic units follow the usage of the Gradstein et al. (2012) and Cohen et al. (2012). Age estimates of boundary positions follow Gradstein
et al. (2012) but are rounded to the nearest whole number (1 Ma) for the pre-Cenomanian, and rounded to one decimal place (100 ka) for the Cenomanian
to Pleistocene interval. Numbered series and stages of the Cambrian are provisional. The Pleistocene is divided into four ages, but only two are shown
here. What is shown as Calabrian is actually three ages: Calabrian from 1.8 to 0.78 Ma, Middle from 0.78 to 0.13 Ma, and Late from 0.13 to 0.01 Ma.
on August 30, 2015gsabulletin.gsapubs.orgDownloaded from
The Geological Society of America Geologic Time Scale
Geological Society of America Bulletin, March/April 2013 261
about…Ephesos and the Plain of the Maiander, if it
be permitted to compare small things with great. And
small these are in comparison, for of the rivers, which
heaped up the soil in those regions none is worthy to
be compared in volume with a single one of the mouths
of the Nile, which has fi ve mouths.”
—Herodotus, likely the world’s fi rst geologist,
fi fth century B.C., in his Histories, 2.10.0-2.
The quest to understand geologic time has
been integral to the geosciences for over 200
years. James Hutton fi rst formally presented the
scientifi c hypothesis that Earth is ancient in a
reading at the Royal Society of Edinburgh on
4 April 1785. He concluded his revolutionary
text, Theory of the Earth, which was based on
his lectures to the Royal Society of Edinburgh
and published in 1788, with the now-famous
and often-quoted sentence concerning the natu-
ral history of the Earth: “The result, therefore,
of our present enquiry is, that we fi nd no ves-
tige of a beginning,—no prospect of an end”
(p. 304). His work in part inspired the great
advances in the geosciences over the following
century, over many parts of the world, and, after
the discovery of radioactivity, prompted the ini-
tial attempts to quantify the age of the planet
Earth and geologic time.
The fi rst attempts by geologists to quantify
the chronostratigraphic time scale and to es-
tablish some bounds on the age of Earth fall
under the category of “hourglass” methods. The
two most important of these involved consid-
erations of the thickness of sedimentary strata
and the salinity of the oceans. Both of these ap-
proaches relied on using estimated rates of geo-
logic processes to establish a more quantitative
time scale. For the fi rst hourglass, the rates of
sediment accumulation were compared to thick-
nesses of sedimentary strata of known chrono-
strati graphic age to compute the duration of
deposition of the rock unit. The second method
relied on understanding the input from streams
to the oceans to compute an overall age for the
oceans. In both cases, these were very rough
approximations of the duration of processes,
because the estimates of rates were inexact.
Importantly, these estimates indicated that the
Earth was much older than was accepted at
the time. This was also true of estimates by
Thomson (see below). Remarkably, during the
second half of the nineteenth century, prior to
the discovery of radioactivity, a number of
workers recognized astronomically forced sedi-
mentary deposits and used them as a means to
calibrate geologic time (reviewed in Hilgen,
2010). The fi rst attempts were built on the astro-
nomical theories for ice ages, and they used
eccen tricity maximums to tune deposits of the
last glaciation but were also tuning Miocene and
Cretaceous sedimentary rocks (Hilgen, 2010).
The discovery of radioactivity, and, more
specifi cally, the relation between radioactive
parent elements and their intermediate and ulti-
mate daughter products through a fundamental
half-life of radioactive decay, was the seminal
event that led to establishing the numerical
ages of geologic materials and ultimately de-
veloping the fi rst chronometric time scale. The
fi rst attempt was by Arthur Holmes (1913) in
his book The Age of the Earth. Holmes (1913,
Chapter X) extensively reviewed the early
methods of geochronology using U as the
parent element. Work at that time showed that
the decay of U produced He as a by-product
and probably had as its ultimate daughter the
element Pb. The half-life of U and thus pro-
duction rates of these elements were roughly
established by the time of Holmes’ (1913)
publication. At the time of Holmes’ work,
it was simply the ratio of U to He or Pb that
was measured—the discovery that elements
had multiple isotopes was reported separately
in the same year (Soddy, 1913; Thomson,
1913). Holmes (1913) reviewed all available
age determinations for U-Pb and U-He that
had bearing on tying numerical age estimates
to meaningful geologic ages. In this effort,
he noted that U-He age estimates were typi-
cally too young, represented minimum ages,
and were most useful for younger (Cenozoic)
rocks. Holmes established the pre-Cenozoic
time scale using a total of fi ve U-Pb determina-
tions. Of these, only three were from rocks of
Phanerozoic age: an age of 340 Ma for the end
of the Carboniferous, 370 Ma for the end of the
Devonian, and 430 Ma for the end of the Ordo-
vician. Ultimately, Holmes used fi ve U-He and
fi ve U-Pb dates to calibrate his geologic time
scale. The ages of other parts of the chrono-
strati graphic time scale not covered by avail-
able age estimates were approximated by using
compilations of sediment thicknesses. Holmes
concluded this chapter by stating (p. 165):
“Most of the available evidence drawn from radio-
active minerals has now been passed in review. As yet
it is a meager record, but, nevertheless, a record brim-
ful of promise. Radioactive minerals, for the geologist,
are clocks wound up at the time of their origin. After
a few years’ preliminary work, we are now confi dent
that the means of reading these time-keepers is in our
possession. Not only can we read them, but if they
have been tampered with and are recording time in-
correctly, we can, in most cases, detect the error and
so safeguard ourselves against false conclusions.”
Besides establishing a rough time scale, this
work was radical in that it expanded the age of
Earth beyond any previous estimate. Up until
that time, the particularly infl uential work of
Lord Kelvin had placed an upper limit of ca.
40 Ma on the age of Earth (Thomson, 1865).
In his fi rst edition of On the Origin of the Spe-
cies by Means of Natural Selection, Charles
Darwin provided a crude estimate of the age
of Earth of several hundred million years
based on both geology and his assumption of
phyletic gradualism. Interestingly, estimates
of the duration of the Phanerozoic incorporat-
ing radiometric ages in the work of Holmes
(1913) at ~550 Ma (inferred from Holmes’ fi g.
17) and Barrell (1917, p. 892) at ~552 Ma are
well within a few percent of the currently ac-
cepted value.
The time scale was greatly refined by
Holmes in the second edition of his book
(Holmes, 1937). At that time he used almost
30 U-He, U-Pb, and Th-Pb age determina-
tions. Because of the recognition of multiple
isotopes, geochronology relied not just on de-
termining the parent-daughter ratio, but in the
cases of the U-Pb and Th-Pb decay systems,
the actual atomic mass and, to some extent,
the isotope ratios of the daughter products
(Holmes, 1937, Chapter V). Holmes continued
to refi ne the time scale over the next 25 years.
Concurrently, analytical methods increased in
their precision while the number of elements
and thus minerals used for isotopic age deter-
minations increased. For example, Kulp (1961)
primarily incorporated K-Ar dates as numerical
estimates for his compilation of the time scale.
A subsequent symposium in honor of Holmes
resulted in a major effort toward developing a
geologic time scale by a broader community
of geoscientists (Harland et al., 1964). In the
symposium volume, the authors reviewed all
of the signifi cant aspects of and developments
with the geologic time scale from numerical
dates, to hourglass methods, and stratigraphic
constraints. In all, over 300 dates were used to
construct a Phanero zoic time scale.
Over the next 20 years, the standardization
of isotopic decay constants by Steiger and Jäger
(1977) and major improvements to the chrono-
strati graphic time scale, fostered by the Inter-
national Union of Geological Sciences (IUGS)
through the International Commission on Stra-
tigraphy (ICS) and its subcommissions, greatly
advanced the geologic time scale. In particular,
the efforts of the ICS established worldwide
standard defi nitions of the relative geologic
time scale. ICS has and continues to establish
global chronostratigraphic standards known as
global boundary stratotype section and points
(GSSPs) (see following discussion). The work
of IUGS and ICS has culminated in more re-
fi ned geologic time scales for the Phanero-
zoic published by Harland et al. (1982), Odin
(1982), and numerous other authors. The initial
GSA time scale (Palmer, 1983) relied heavily
on the contributions by these authors. A major
on August 30, 2015gsabulletin.gsapubs.orgDownloaded from
Walker et al.
262 Geological Society of America Bulletin, March/April 2013
treatment, including the statistical assessment
of uncertainties of the estimated boundary ages,
was presented by Harland et al. (1989). Con-
siderable progress was made over the next 15
years, with the next major and comprehensive
revision to the geologic time scale published
by Gradstein et al. (2004). This publication
fully integrated the advances made in global
stratigraphy through the work of the ICS, in-
cluding astrochronology, and took advantage
of considerably more precise radioisotope
geo chronol ogy including 40Ar/39Ar step heat-
ing and single-crystal laser methods as well as
U-Pb zircon dating methods.
History of the Geological Society of America
Geologic Time Scale
Besides being the 125th anniversary of the
Geological Society of America, 2013 marks
the 30th anniversary of the Geological Soci-
ety of America Geologic Time Scale, the 50th
anniversary of the publication of the Vine-
Matthews-Morley-Larochelle hypothesis that
marine magnetic anomaly patterns adjacent to
mid-ocean ridges were a record of the polarity
reversal history of the geomagnetic fi eld and
thus that the ridges were sites of ocean-fl oor
spreading (Vine and Matthews, 1963), the 100th
anniversary of the fi rst geologic time scale pre-
sented by Holmes (1913), and the 225th anni-
versary of Hutton’s Theory of the Earth (1788).
The GSA time scale grew out of the Decade
of North America Geology (DNAG) project
(Palmer, 1983), which had as its goal a syn-
thesis of the geology then known of the North
American continent. Before that time, geologic
information on North America had been scat-
tered through the literature. With a few excep-
tions (e.g., Eardley, 1951; King, 1959), little
in the way of comprehensive summaries of the
geology of North America existed. A necessary
prerequisite for discussing the geology of the
continent was a common chronostratigraphic
vocabulary and internally consistent sense of
both the relative and numerical ages of geo-
logic and evolutionary events. This provided
ages of specifi c stratigraphic units addressed in
the multivolume compilation involving scores
of authors that resulted from the DNAG proj-
ect. The DNAG project was the fi rst detailed
synthesis of North American geology follow-
ing widespread acceptance of plate-tectonic
theory and provided a sense of the evolution
of the continent in the context of global events.
Before the DNAG compilation could proceed
very far, a uniform geologic time scale had to
be adopted, and it was the fi rst major product
of the DNAG initiative. This inevitably led to
advancement of a version of the geologic time
scale that was North American–centered but
with more far-reaching consequences. With
progress of the DNAG project and considerable
changes in chronostratigraphic nomenclature
taking place internationally, refi nements of the
time scale were inevitable, and updated versions
were published as GSA’s 1999 Geologic Time
Scale (Palmer and Geissman, 1999) and 2009
Geologic Time Scale (Walker and Geissman,
2009). What set the original DNAG time scale
apart from earlier versions of time scales pub-
lished in textbooks and summary papers was
the overt application of integrated stratigraphic
and magnetostratigraphic data sources (includ-
ing both relative and numerical age dating tech-
niques used to assemble the time scale) and the
style of presentation.
The rationale behind and history of the work
to develop the fi rst Geological Society of Amer-
ica Geologic Time Scale itself were described
by Palmer (1983) and Walker and Geissman
(2009). Allison (“Pete”) Palmer was the Cen-
tennial Science Program Coordinator for GSA’s
DNAG project, and was charged with compil-
ing the results of the efforts of the Time Scale
Advisory Committee, consisting of Z.E. Peter-
man, J.E. Harrison, R.L. Armstrong, and W.A.
Berggren. Pete Palmer had a clear passion for
the work and devoted considerable energy to the
project. An innovative contribution was the at-
tempt to organize the time scale onto a single
8.5 by 11 inch sheet of paper using the “tools
of the trade” in those days—Mylar, zipatone
letters, a Leroy lettering machine, and a lot of
patience. His efforts resulted in a unique lay-
out for the time scale, in which each era of the
Phanerozoic, and all of the Precambrian, was
given identical column length. Scaling of the
magnetic polarity time scale to fi t this format
required numerous trials. When Jim Clark, the
director of publications for GSA, suggested that
the geologic time scale should also be published
as a pocket wallet card, the effort became even
more challenging. A key concern by both the ad-
visory committee and Palmer was whether the
time scale should be North American centric or
global. The advisory committee and Palmer rec-
ognized the likely pitfalls and hurdles associated
with trying to develop a global time scale and
set as a goal trying to develop a “common vo-
cabulary” for North America, still a suffi ciently
great challenge. In discussions with Palmer
(2012, personal commun.), he emphasized the
diffi culties with defi ning the base of the Ordovi-
cian (which has more recently been solved with
the defi nition of an Ordovician GSSP). The fi -
nal product (Palmer, 1983) served the intended
purpose well. “There were no naysayers, and no
major disagreements,” stated Palmer. “The basic
subdivisions were all accepted.”
RECENT ADVANCES AND THEIR
IMPACT ON THE GEOLOGIC
TIME SCALE
Here, we review some of the key advances
that have led to changes and refi nements of the
geologic time scale. These include, of course,
changes in stratigraphic and geochronologic
approaches that are at the heart of a combined
chronometric/chronostratigraphic time scale,
but also astrochronology, which is revolution-
izing the time scale effort, chemostratigraphy
and related rock magnetic stratigraphy, and the
geomagnetic polarity time scale (see Table 1 for
brief descriptions of these methods).
Advances in Stratigraphy—The
International Commission on Stratigraphy
Leading the way in the advancement of
chronostratigraphic information is the Inter-
national Commission on Stratigraphy, including
the many subcommissions formed by the ICS.
These groups actively promote the acquisi-
tion and dissemination of information vital to
making informed decisions about stratigraphic
boundary positions and numerical calibrations
of those positions. A major goal of the ICS is
to develop an unambiguous, globally applica-
ble nomenclature for geologic time units and
their chronostratigraphic equivalents, a com-
mon language in which concepts are identical
for all localities across the globe. This effort is
part of a larger project to develop unambiguous
chronostratigraphic units for the entire geologic
time scale (Gradstein et al., 2004, 2012; Ogg
et al., 2008; Cohen et al., 2012). The main driv-
ing forces for the refi nements and restructuring
to the chronostratigraphic side of the geologic
time scale are changes in philosophy about
how stratigraphic units are defi ned, as well as
high-resolution studies utilizing chronostrati-
graphic proxies, including biostratigraphy,
chemostratigraphy, magnetostratigraphy, rock
magnetic stratigraphy, and astrochronology/
orbital tuning.
Over the past several decades, different strati-
graphic philosophies have been applied to defi -
nitions of chronostratigraphic units. Defi nitions
in many places around the world were com-
monly based on the unit-stratotype concept, in
which a type section serves as the standard of
reference for the defi nition and characterization
of a unit (Salvador, 1994). The lower and upper
boundaries of a unit are normally specifi ed by
reference to a type section. Beginning in the
1970s, the concept of a boundary-stratotype was
promoted. Under this concept, only the base of a
chronostratigraphic unit is formally defi ned, and
it is marked by a point in strata (Salvador, 1994),
on August 30, 2015gsabulletin.gsapubs.orgDownloaded from
The Geological Society of America Geologic Time Scale
Geological Society of America Bulletin, March/April 2013 263
which is known as a GSSP, or global boundary
stratotype section and point, more popularly re-
ferred to as a “golden spike.”
Originally, the boundaries of most geologic
time units were drawn at positions in the strati-
graphic record where some sizable biotic change,
widespread geologic event such as a glacial
advance or retreat, or an orogenic episode was
recognized. Some of the formal nomen clature
of the geologic time scale refl ects this practice:
“Phanerozoic” means “visible life,” Paleozoic,
Mesozoic, and Cenozoic mean “ancient life,”
“middle life,” and “recent life,” respectively. In-
formal terms applied to geologic time units, such
as the “age of bryozoans” for the Ordovician,
the “age of coal swamps” for the Carboniferous,
the “age of dinosaurs” for the Mesozoic, and the
“age of mammals” for the Cenozoic, reinforce
this approach to characterizing the progression
of geologic time. This broad-brush view may
have some merits, but to a large extent, the ap-
parent distinctions between strata contained
in superjacent chronostratigraphic units have
been enhanced by evolutionary overturns (ex-
tinctions, followed by recoveries and changing
patterns of ecologic dominance) and unconfor-
mities of varying scale. Boundaries drawn at
such horizons are almost certain to be in strati-
graphic gaps. In the past, chronostratigraphic
units based on the unit-stratotype concept were
often drawn at previously unrecognized gaps in
the stratigraphic record. To be useful, a globally
synchronous boundary must be drawn at a hori-
zon in a continuous succession of strata to en-
sure that there are no gaps in the succession. A
boundary-stratotype–based chronostratigraphic
unit is defi ned only at the base, so the “top” is
automatically defi ned by the base of the overly-
ing unit, and there can be no gap in between.
GSSPs are defi ned according to the boundary-
stratotype concept. By defi nition, every GSSP
marks the base of a chronostratigraphic unit.
Also by defi nition, the chronostratigraphic
unit subjacent to each GSSP includes all strata
known and unknown up to that horizon, so there
can be no gap between adjacent chronostrati-
graphic units.
High-resolution stratigraphic studies today,
often enhanced with continuous drill cores,
are the norm, so more detailed information is
available now than when early versions of the
geologic time scale were assembled. The result-
ing details are not always clear, and unambigu-
ous sets of data in some cases have shown that
geoscientists were sometimes mistaken in their
assumptions of isochronous events. The fi rst
appearance of trilobites was once used through
much of the world as the marker for the base
of the Phanerozoic. Today, we recognize that
the appearance of trilobites on separate paleo-
conti nents was not synchronous, making such
an inferred singular “event” unsuitable on
its own as a defi nitive correlation tool. High-
resolution chronostratigraphic studies using
multiple methods in combination now give us
the ability to pinpoint an exact datum in strata
where a boundary can be drawn. For example,
the base of the Cenozoic coincides with the base
of an iridium-bearing clay layer, as fi rst demon-
strated at Gubbio, Italy (Alvarez et al., 1980).
On a global scale, numerous other stratigraphic
tools (among them, the biostratigraphic ranges
of nannoplankton and foraminiferans, spores
and pollen, ratios of stable isotopes of carbon
and oxygen, and magnetic polarity stratigraphy)
provide information that adds to the anomalous
concentration of iridium at that single datum to
help defi ne its singular position. Geochronology
also can be used to identify the base of the Ceno-
zoic with great precision (ca. 66.0 Ma), adding
to the broad range of techniques that can be used
to identify the boundary horizon globally.
Most GSSPs in the Phanerozoic have been
defi ned to coincide with an evolutionary event,
normally the evolutionary fi rst appearance of a
recognizable guide fossil. For at least one chrono-
stratigraphic unit, this approach is impractical,
in part because some available time proxies
exceed biostratigraphy in their ability to pro-
vide high-resolution stratigraphic information.
The base of the Holocene Series/Epoch, for
example, has been defi ned at a point refl ecting
a distinct switchover in deuterium content of
glacial ice in Greenland. That point essentially
coincides with the onset of warming at the end
of the Younger Dryas and a change in thickness
of glacial ice layers, refl ecting increased dust
content associated with the warming interval.
Chemostratigraphy, particularly using δ13C, pro-
vides high-resolution stratigraphic information
for parts of the Paleozoic that have been chal-
lenging to defi ne on the basis of biostratigraphic
evidence alone. To date, stable isotope data have
not been used as a primary tool for defi ning a
chronostratigraphic unit, but biostratigraphi-
cally calibrated isotopic excursions, particularly
in δ13C, have some advantages in successions
where the biostratigraphic data are inconclusive
at a detailed scale (e.g., Maloof et al., 2010;
Korte and Kozur, 2010). As we achieve a more
highly calibrated record, assuming global syn-
chroneity of the Carbon record, for example,
may not remain valid. Since the completeness
of sections is variable, it may be that no single
proxy will resolve all issues making the use and
integration of multiple proxies required.
The application of the GSSP approach has
helped stabilize the geologic time scale by
adoption of chronostratigraphic nomenclature
ratifi ed by specialists the world over. Changes to
the time scale resulting from these advances are
numerous and some time periods have been dra-
matically restructured through formal decisions
on boundary positions. The Cambrian System/
Period, for example, now has a ratifi ed base
that is stratigraphically well below its position
TABLE 1. TIME DETERMINATION METHODS
Methods Definition and application
Astrochronology Study of cyclical variation in properties present in sedimentary rock sequences interpreted to represent a response to Earth orbital
parameters deriving from interaction with the Sun, Moon, and other planets. Variations can be expressed in physical sedimentary
patterns (e.g., bed thicknesses) or chemical characteristics. Can be used for establishing durations in the Cenozoic and in older time
intervals if calibrated by geochronologic results and for direct dating in rocks younger than ca. 35 Ma.
Chemostratigraphy Study of chemical and isotopic variations in sedimentary sequences interpreted to reflect changes in seawater chemistry or catastrophic
events (e.g., Ir anomaly at the end of the Cretaceous). Detailed variations in chemical signals may reflect astronomical forcing
and can be used as a dating tool when signal changes at a high rate (Cenozoic Os and Sr). Can also be used for relative time by
correlation of major unique signatures.
Geochronology Study of radioisotopes and their decay products. Time measured by the ratio of radioisotopes to decay products using a decay
constant. This is a numerical dating method.
Magnetic polarity
stratigraphy Study of the polarity record of Earth’s magnetic field as reflected in the paleomagnetic signatures of stratified rocks. This gives relative
ages and allows for correlation of sequences. It can approximate a numerical method if one or, ideally, many parts of a sequence can
be tied to a numerical age.
Rock magnetic
stratigraphy Study of the variations in the magnetic mineralogy (concentration and mineral phases) assumed to be originally deposited with the
sedimentary sequence. At a fine-scale examination (e.g., bed or even lamina), this method can be a component of astrochronology.
Stratigraphy Study of stratigraphic successions by using the physical arrangement of layers and sequences (lithostratigraphy) as well as the fossil
content (biostratigraphy). This is a relative dating and correlation method.
on August 30, 2015gsabulletin.gsapubs.orgDownloaded from
Walker et al.
264 Geological Society of America Bulletin, March/April 2013
in 1983, and the ratifi ed base of the Ordovician
System, which marks the top of the Cambrian,
is above its position in 1983. The 1983 position
of the Cambrian base, following Harland et al.
(1982) was at the base of the Tommotian Stage
as used in Siberia and, at that time. Today, that
position is within provisional Stage 2 of the
Cambrian System and estimated to be about
525 Ma (570 Ma by 1983 standards). The hori-
zon is well within what was in 1983 considered
to be Proterozoic strata. The base of the Ordovi-
cian, after formal defi nition, is at a horizon near
485 Ma (504 by 1983 standards).
Formal decisions on chronostratigraphic
boundaries have in some instances led to
changes in the way we subdivide systems/
periods . Largely as a result of the addition of a
thick stratigraphic section below its traditional
base, the Cambrian is now subdivided into four
series/epochs rather than the traditional three,
and the resulting names for the epochs are de-
rived from localities, not relative stratigraphic
position. The Silurian Period is also subdivided
into four epochs, the names of which are de-
rived from localities. The Permian Period is
subdivided into three epochs, and their names
are likewise derived from localities rather than
relative stratigraphic position. The Cenozoic
Era has been considerably reorganized. It now
consists of three periods, Paleogene, Neogene,
and Quaternary. Strata traditionally assigned to
the uppermost Pliocene Series/Epoch (Neogene
System/Period) are now assigned to the Gela-
sian Stage/Age, which belongs in the Pleisto-
cene Series/Epoch (Quaternary System/Period).
As a result, the Quaternary is 43% longer to-
day than it was prior to addition of the Gela-
sian Stage/Age. Importantly, the base of the
Holo cene Series/Epoch has been defi ned, and
its beginning is dated at 11,700 yr b2k (before
A.D. 2000). Terms such as Tertiary, Recent,
and Anthropocene have been dropped from the
geologic and chronostratigraphic vocabulary as
they no longer have any internationally sanc-
tioned standing.
Advances in Geochronology and the
EARTHTIME Effort
The past decade has seen dramatic advances
in high-precision geochronology. These ad-
vances have been used to develop time lines for
Earth history and deconvolve complex tectonic
and magmatic events. Importantly, the applica-
tion of high-precision geochronology to sedi-
mentary sequences has allowed researchers to
determine the age, duration, and possible syn-
chroneity of global events such as extinctions,
rates of biological evolution and changes in
biodiversity, and the age and duration of stable
isotope anomalies that refl ect major changes in
seawater and atmospheric chemistry. This
mainly results from dramatic increases in the
precision of radioisotopic dates, especially
using 40Ar/39Ar (feldspars) and U-Pb (zircon)
methods. These increases have been stimulated
by new laboratory methods, new generations of
mass spectrometers, and scientifi c projects that
demand the highest possible precision. The U-Pb
method can be applied to rocks from as young
as 600 ka, to as old as the oldest known min-
eral grains on the planet (ca. 4.4 Ga zircon), to
meteorites. The 40Ar/39Ar method is mainly used
for rocks that range in age from younger than
600 ka through the Paleozoic, and even in some
cases Archean rocks and meteorites. However,
40Ar/39Ar systematics are typically disturbed by
metamorphism in rocks exposed to greenschist
and higher grades of metamorphism. For rocks
younger than 600 ka, 40Ar/39Ar and U-series
geochronology are best applied. The time in-
terval refl ecting recent landscape evolution and
erosion is best evaluated using several different
terrestrial cosmogenic nuclides.
As precisions improved (e.g., <0.1% for
U-Pb zircon methods), it was apparent that there
were both intertechnique and interlaboratory
systematic errors for the U-Pb and 40Ar/39Ar
methods. Thus the EARTHTIME project (http://
www.earth-time.org/ and European sister orga-
nization http://earthtime-eu.eu/) was initiated
to develop a community-driven approach to
the calibration of Earth history using 40Ar/39Ar
and U-Pb methods. The project organized all
the major geochronology laboratories in the
world in a collaborative effort to eliminate both
inter laboratory and intertechnique biases and
to critically assess both accuracy and precision
associated with numerical ages. The ultimate
goal is to calibrate Earth history, and the effort
has already produced remarkable cooperation
among geochronologists and the rest of the earth
science community, especially paleontologists,
astrochronologists, and stratigraphers. This has
also spawned a new generation of students who
range from conversant to full practitioners of
radioisotope geochronology and paleontology,
stratigraphy, and astrochronology. A critical ele-
ment of this globally collaborative approach is
the need for transparency for all laboratory tech-
niques, from sample preparation to data acqui-
sition and data reduction to data reporting and
archiving.
The fi rst step for the U-Pb community was to
eliminate a large source of interlaboratory bias
caused by the fact that each laboratory used a
different isotopic tracer. Enriched tracers are
added to U-bearing accessory minerals prior
to dissolution to allow the U/Pb ratio of the
mineral to be determined. Tracers used by dif-
ferent laboratories when EARTHTIME began
included both 208Pb and 205Pb that were mixed
with 235U or 233U + 235U or 233U + 236U. Two iso-
topes of U allow mass-dependent fractionation
to be evaluated during the run, greatly reduc-
ing one of the biggest sources of uncertainty
in U isotopic analyses. In addition, there are a
few tracers containing 233U-235U-202Pb-205Pb that
allow correction for fractionation of the Pb iso-
topes as well. It was decided that the community
needed a single mixed tracer for all laboratories
interested in the EARTHTIME goal (producing
the highest-precision geochronology to estab-
lish a calibrated time line for Earth history) and
one was mixed, calibrated, and sent to several
laboratories (Condon et al., 2007).
At about the same time, U-Pb zircon geo-
chronol ogy was forever changed by the develop-
ment of a method that eliminates open-system
behavior (usually Pb loss) from zircon. The
method was developed by James Mattinson
(Mattinson, 2005) and involves high-tempera-
ture annealing of grains followed by partial dis-
solution. Mattinson was able to show that the
high-U parts of the grains that are most suscep-
tible to Pb loss can be preferentially removed,
leaving relatively low-U domains with no evi-
dence of Pb loss.
EARTHTIME workers did a series of blind
tests where unknown zircons and synthetic U-Pb
solutions were sent to a number of laboratories
with the EARTHTIME tracer. The fi rst blind test
was done without the tracer, and dispersion was
close to 1%. The second test was done using the
EARTHTIME tracer, and the synthetic solutions
indeed greatly decreased interlaboratory vari-
ability, and at least six laboratories now agree
at the 1‰ level. Finally, the community seeks to
have common open-source data acquisition and
reduction protocols that can evolve as new ideas
are developed. The same approach can be and is
being used for all techniques from 14C to terres-
trial cosmogenic nuclides (http://www.physics.
purdue.edu/primelab/CronusProject/cronus/).
The 40Ar/39Ar community is working in par-
allel on developing neutron fl uence monitors,
including independent assessment of their age,
other age standards, and data acquisition of fl u-
ence monitors. This community also has run two
blind tests. The fi rst test, involving fi ve samples,
indicated considerable dispersion among labora-
tories but no systematic bias. In the second test,
all fi ve samples were sent for one irradiation and
then distributed, eliminating differences in reac-
tors and irradiation protocols. The results were
similar to the fi rst with ~2% dispersion among
laboratories, and this is much greater than the
reported analytical precisions (mostly 0.1%–
0.5%). There are several potential explanations,
including differences in the data acquisition and
on August 30, 2015gsabulletin.gsapubs.orgDownloaded from
The Geological Society of America Geologic Time Scale
Geological Society of America Bulletin, March/April 2013 265
reduction protocols in each laboratory (currently
there is not a common platform), nonlinearity of
the mass spectrometer source/detector systems,
and different systems for cleaning sample gas
using getters. A new approach to solving this
problem is under way. A pipette system fi lled
with gases of a known isotopic composition will
travel to participating laboratories to eliminate is-
sues associated with sample heterogeneity. There
will be three gasses in the pipette experiment;
one will be air, and the other two will be aliquots
of a large natural sample of biotite irradiated dif-
ferent times such that the ratio of 40Ar*/39Ar is
~23.6 between the two experiments. An impor-
tant aspect of the pipette experiment is that each
canister will have three calibrated pipette vol-
umes of 0.1, 0.2, and 0.4 cm3, allowing for vary-
ing the amount of gas delivered as 0.1, 0.2, 0.3
(0.1+0.2), 0.4, 0.5 (0.1+0.4), 0.6 (0.2+0.4), and
0.7 cc. Furthermore, it will be possible to mix
these gases with the same volume increments
to produce intermediate values. This approach
should better elucidate the sources of interlabora-
tory dispersion. We are highly optimistic that
within a few years, the 40Ar/39Ar community will
have eliminated interlaboratory bias and have in-
ternal errors of ~0.3%, if not lower.
As EARTHTIME community members have
attempted to improve precisions of radioiso-
topic dates, there has been a much deeper ap-
preciation of the ways in which to quantify most
sources of uncertainty and to fully propagate
associated errors. This in turn has led to new
protocols for data acquisition, reduction, and
archiving through open-source software (Bow-
ring et al., 2011; McLean et al., 2011), result-
ing in more accurate uncertainties in calculated
U-Pb dates. Archiving of data will be achieved
through a collaboration with EarthChem (at
http://www.geochron.org), where data may be
uploaded directly from the software and will al-
low future recalculation of dates as “constants”
such as the 238U/235U ratio in zircons (e.g., Hiess
et al., 2012), U and K decay constants, and tracer
calibration changes due to new measurements
in any laboratory while preserving the original
uploaded data. This will allow any time scale
to be updated using published and recalculated
dates. The 40Ar/39Ar community is also working
toward a common platform for data reduction
and is linked to EarthChem for archiving of data
through current data reduction software.
Advances in the Geomagnetic Polarity
Time Scale—Developments and Integration
with the Geologic Time Scale
“The critical thing at that meeting was that I met
Brent Dalrymple for the fi rst time. He told me in pri-
vate discussion between sessions, ‘We think that we’ve
sharpened up the polarity reversal scale a bit, but in
particular we’ve defi ned a new event—the Jaramillo
event.’ I realized immediately that with that new time
scale, the Juan de Fuca Ridge could be interpreted
in terms of a constant spreading rate. And that was
fantastic, because we realized that the record was
more clearly written than we had anticipated. Now
we had evidence of constant spreading; that was very
important.
“I realized at once, having poured over the prob-
lem for so long and so recently, that with this revised
time scale it would be possible to interpret the Juan de
Fuca anomaly sequence with an essentially constant
rate of spreading. To me, at that instant, it was all over,
bar the shouting.”
—Frederick J. Vine, recalling his meeting of
Brent Dalrymple at the 1965 GSA Annual
Meeting, November, Kansas City, in Glen
(1982) and Oreskes (2001), respectively.
It has been 50 years since the publication
of the seminal contribution by Fred Vine and
Drummond Matthews (Vine and Matthews,
1963) in which marine magnetic anomaly pat-
terns were interpreted in the context of seafl oor
spreading and mid-ocean ridges (the Vine-
Matthews-Morley-Larochelle hypothesis). It
was the recognition that Earth’s magnetic fi eld
was capable of reversing its polarity, and in
fact had done so numerous times, that allowed
their hypothesis to be proffered. This was at a
point in the early stages in the history of the
development of the geomagnetic polarity time
scale and magnetic polarity stratigraphy, as the
“order ing of sedimentary or igneous rock strata
into intervals characterized by the direction of
magnetization of the rocks, being either in the
direction of the present Earth’s fi eld (normal
polarity) or 180° from the present fi eld (reverse
polarity)” (Opdyke and Channell, 1996). It was
also at a point soon after the general acceptance
of the fact that the geomagnetic fi eld was capa-
ble of reversing its polarity—a concept that for
decades met with considerable argument and
debate in the scientifi c community. The fi rst
report of a magnetization in a rock (the natural
remanent magnetization, NRM) with a direc-
tion reverse to that of the present geomagnetic
fi eld of Earth was that of Brunhes (1906), who
examined Pliocene lava fl ows from central
France. Several other studies over the ensuing
decades reported magnetizations reverse to that
of the present fi eld, the most notable of which
is that by Matuyama (1929), who suggested the
possibility that the polarity of the geomagnetic
fi eld was age dependent. With the development
of magnetometers with suffi cient sensitivity to
measure the NRM of sedimentary rocks, mag-
netic polarity stratigraphy began to blossom in
the late 1940s and early 1950s. The earliest such
study in North America was that by Torreson
et al. (1949), which demonstrated that magneti-
zations of reverse polarity were retained in rocks
as old as the early Mesozoic. Numerous studies
further enforced the hypothesis that Earth’s fi eld
had reversed its polarity in the past.
These emerging data served to enforce the
need to settle the controversy concerning the pos-
sibility that the fi eld was capable of reversing
polarity, as workers recognized the potential of
Earth’s polarity history for establishing an inde-
pendent scale of geologic time. In support of the
fi eld reversal hypothesis, there was a growing
body of data showing positive “baked contact”
tests, where the magnetization in a host rock im-
mediately adjacent to an intrusive igneous rock,
or below a lava fl ow, was similar if not identical
to that in the igneous rock, and thus of the same
polarity. In his classic early text on paleomag-
netism, Irving (1964) described the ideal baked
contact test (his fi g. 7.22, p. 172) that, once and
for all, would convincingly demonstrate that
reverse polarity magnetizations in rocks were
truly a geomagnetic fi eld phenomena, rather
than some form of “self-reversal” characteristic
of a complicated magnetic mineralogy that is not
common to most rocks. As Opdyke and Chan-
nel (1996) noted, the geomagnetic fi eld reversal
hypothesis was gaining considerable support by
this time. First, all rocks of late Pleistocene and
Holocene age were of normal polarity. Second,
reverse polarity magnetizations were common
to rocks of early Pleistocene age. Third, mag-
netizations with directions that were “transi-
tional” between normal and reverse polarity
were identifi ed in both sedimentary and igne-
ous rocks; their presence would be consistent
with changes in polarity states occasionally be-
ing recorded. Fourth, self-reversal mechanisms
were identifi ed in laboratory experiments, but
in only a few types of rocks; the overwhelming
majority of experiments demonstrated that most
rocks acquired a thermoremanent magnetization
that replicated the ambient fi eld. Finally, many
baked contact tests, including some involving
the nearly ideal fi eld conditions postulated by
Irving (1964), provided strong support for dual
polarities of the geomagnetic fi eld.
As described in The Road to Jaramillo (Glen,
1982), tremendous competition and excitement
raged between teams of geochronologists/paleo-
magnetists from laboratories at the U.S. Geo-
logical Survey at Menlo Park and University
of California at Berkeley and their Australian
counterparts at the Australian National Uni-
versity. The important geochronologic method
of potassium/argon age dating was being per-
fected, and the “race” was on to establish a more
refi ned polarity time scale, based on isotopic
age determinations of the very rocks examined
for polarity information. By 1963, the combined
geochronologic and magnetic polarity data were
suffi ciently compelling in their internal consis-
tency to allow Vine and Matthews to formulate
on August 30, 2015gsabulletin.gsapubs.orgDownloaded from
Walker et al.
266 Geological Society of America Bulletin, March/April 2013
and have accepted for publication their now fa-
mous explanation of mid-ocean ridges and their
importance. In the context of both the geomag-
netic polarity time scale and global tectonics,
the Vine and Matthews (and Morley/Larochelle)
hypothesis of seafl oor spreading was certainly
not immediately embraced by the broad commu-
nity. Skeptics noted the fi ne structure of marine
magnetic anomaly patterns adjacent to crests of
mid-ocean ridges and pointed out inconsisten-
cies between anomaly patterns and the current,
yet quickly evolving, polarity time scale. Much
of that skepticism disappeared following the
Doell and Dalrymple (1966) paper in Science,
in which they demonstrated that the last ~1 m.y.
of Earth history was not of entirely normal
polarity, but rather consisted of some 700 ka
of normal polarity, followed by an interval of
~100 ka of reverse polarity, an ~150 ka interval
back to normal polarity, and a several 100 ka
interval of reverse polarity, which character-
ized what was then the early Pleistocene. They
named the short time interval of normal polar-
ity the Jaramillo “event,” after Jaramillo Creek
in the Valle Grande of the Jemez Mountains,
north-central New Mexico, and also described
the longer intervals of constant polarity as mag-
netic epochs (e.g., Brunhes normal, Matuyama
reverse, Gauss normal, and Gilbert reverse). As
noted previously herein, Fred Vine fi rst learned
of this new observation in November 1965. By
spring 1967, the hypothesis of seafl oor spread-
ing, and its many implications, had risen in
status to the theory of plate tectonics.
At the time of publication of the 1983 Geo-
logic Time Scale (Palmer, 1983), our under-
standing of the geomagnetic polarity time scale
had evolved by incremental refi nements from
a chronology of the last few million years of
Earth history to an increasingly robust defi -
nition of Earth’s polarity history back to the
Middle Juras sic. How this happened in consid-
erably less than 20 years involved additional
high-quality radioisotopic dates obtained from
igneous rock sequences with well-defi ned po-
larity records, improved magnetic polarity stra-
tigraphy records, including those from deep-sea
sediments, determination of the approximate
ages of boundaries between some specifi c po-
larity epochs, and a global integration of marine
magnetic anomaly data into an interpretable rec-
ord of polarity changes, eventually back to the
Middle Jurassic. These efforts were summarized
by Berggren et al. (1985a, 1985b) and Kent and
Gradstein (1986). Having several key numeri-
cal age estimates of polarity boundaries allowed
for further refi nements to the magnetic polar-
ity time scale based on the marine magnetic
anomaly record (Cande and Kent, 1992, 1995)
and better recognition of the fact that seafl oor
spreading rates were not constant among ocean
lithosphere plates and that they have not been
constant since the early Mesozoic.
Since 1983, the geomagnetic polarity time
scale has been extended to the base of the Tri-
assic and in fact the very latest Permian using
magnetic polarity stratigraphy records from
continental and marine sedimentary rocks, in
combination with high-resolution geochrono-
logic data, biostratigraphy, and cyclostratig-
raphy, where possible. It may be possible to
extend an accurate, chronologically defi ned
polarity time scale into the latest Mississippian,
prior to the Permian-Carboniferous Reverse
Superchron. This superchron is perhaps the
singular most unusual feature of the geomag-
netic fi eld during the entire Phanerozoic, as we
have little evidence to contradict the likelihood
that the fi eld was exclusively of reverse polar-
ity for some 55 Ma. The termination of the re-
verse superchron, originally named the Kiaman
Magnetic Interval by Irving and Parry (1963), is
close to the Wordian-Capitanian boundary (ca.
265 Ma). The base of the superchron, however,
is less well known. Opdyke et al. (2000) argued
that it might lie in lowermost Pennsylvanian
strata. For most of the remainder of the Paleo-
zoic, the magnetic polarity record is relatively
ill defi ned, as noted by several workers and sum-
marized by Opdyke and Channell (1996). The
exception to this statement is an interval (super-
chron) of reverse polarity in the Early to Middle
Ordovician that may be some 30 Ma in duration.
Pavlov and Gallet (2005) proposed the name of
“Moyero” for the superchron, based on the lo-
cation in Siberia where a fi rst complete record
of the interval was identifi ed. They also noted
that the Middle Cambrian was a period of high
polarity reversal frequency. Subsequent work by
Pavlov et al. (2012) has further documented the
chronostratigraphic framework for the Moyero
superchron.
Two important time intervals in Earth history
for which considerable effort has been made to
understand and defi ne, at very high resolution,
the magnetic polarity time scale are the Late
Triassic to earliest Jurassic and the late Middle
Permian to earliest Triassic. Both of these time
intervals include major extinction events and
the development of large igneous provinces, the
Central Atlantic magmatic province, and Sibe-
rian fl ood basalts, respectively.
The magnetic polarity record of the Triassic,
and particularly the Late Triassic, has evolved
considerably since early work by Keith Run-
corn and colleagues in the 1950s. The seminal
work by Paul Olsen and Dennis Kent and col-
leagues on the continuously cored Upper Trias-
sic sequence of the Newark Basin (Kent et al.,
1995; Olsen et al., 1996a, 1996b) provided a
high-resolution, astrochronologically tuned po-
larity record for the ca. 25+ Ma. time interval
obtained, which has been incorporated into the
GSA geologic time scale. This advance permit-
ted far more robust correlations of magnetic
polarity data obtained from marine sections that
preserved more complete records of evolution-
ary change during the latest Triassic to earli-
est Jurassic transition. This time interval (late
Rhaetian to earliest Hettangian) is dominantly
of normal polarity. Importantly, as summa-
rized by Lucas et al. (2011), individual sections
across the boundary have revealed between two
and four reverse polarity zones inferred to be of
short duration, and thus microzones. If possible,
the precise correlations of these microzones, in
marine and continental sections, across the Tri-
assic-Jurassic boundary may ultimately prove
invaluable in understanding the timing of end-
Triassic vertebrate extinctions and the evolution
of Jurassic recovery faunas.
Several recently obtained magnetic polarity
stratigraphy records from both marine and ter-
restrial strata across the Permian-Triassic bound-
ary interval show that both the end-Permian
ecological crisis as well as the conodont-
calibrated biostratigraphic Permian-Triassic
boundary both followed a key polarity reversal
from a relatively short interval (subchron) of re-
verse polarity to a considerably longer interval
(chron) of normal polarity (Li and Wang, 1989;
Scholger et al., 2000; Szurlies et al., 2003). An
excellent example of the importance of inte-
grating biostratigraphic, cyclostratigraphic, and
magnetic polarity stratigraphic records is the
principally continental Permian (Rotliegend
and Zechstein Groups) and immediately over-
lying epicontinental Triassic (Buntsandstein
Group) strata from Western and Central Europe,
which have yielded high-quality magnetic po-
larity stratigraphic records. In combination with
cyclo strati graphic records, Szurlies et al. (2003,
2012) estimated that the normal polarity chron
containing both the end-Permian crisis and the
biostratigraphic Permian-Triassic boundary was
~0.7 Ma in duration.
Chemostratigraphy
Chemostratigraphy is the study of variations
in the primary chemical and isotopic composi-
tions of sedimentary rocks with time. A vari-
ant of chemostratigraphy is rock magnetic
stratigraphy, where variations in the concen-
tration and/or mineralogy of magnetic phases
are determined in sedimentary sequences (e.g.,
Evans and Heller, 2003; Kodama et al., 2010).
Ideally , in chemostratigraphic and rock mag-
netic stratigraphic approaches, it is assumed
that the chemical/rock magnetic signals pre-
on August 30, 2015gsabulletin.gsapubs.orgDownloaded from
The Geological Society of America Geologic Time Scale
Geological Society of America Bulletin, March/April 2013 267
served within sedimentary rocks provide a
proxy for global seawater chemistry and/or en-
vironmental conditions at the time of, or shortly
after, depo si tion. However, postdepositional
processes can overprint primary signals, and
much care must be taken to avoid interpretation
of signals that have been modifi ed by postdepo-
sitional processes as primary. If one assumes
that the variation in seawater chemistry is a
global and homogeneous signal, then one of the
main uses of chemostratigraphy is to correlate
stratigraphic variability in chemistry between
packages of sedimentary rocks across a basin
and/or around the globe. The relative tempo-
ral framework provided by chemostratigraphic
correlation is enhanced using biostratigraphy,
magnetostratigraphy, floating astrochronol-
ogy (see following section), and calibration of
chemical signals with numerical time using
either well-established boundaries (e.g., extinc-
tions, isotopic anomalies) or accurately dated
volcanic rocks. Most recently defi ned GSSPs
are at horizons close to some well-established
chemostratigraphic shift that can be correlated
between sections (Gradstein et al., 2012). In
strata that are devoid of recognizable guide
fossils, chemostratigraphic correlation has
been essential to unraveling the chronology of
Earth history. Study of the Neoproterozoic has
benefi ted greatly from applying stable isotope
chemostratigraphy to chronostratigraphic corre-
lation (e.g., Knoll and Walter, 1992; Halverson
et al., 2005).
A major complication to detailed chemostrati-
graphic correlations across large distances is
that it is unlikely that multiple depositional
areas capture the same complete and continu-
ous record of sediment accumulation. In addi-
tion, accumulation rates between locations often
vary signifi cantly. In detail, these factors result
in different structures of chemical variability, re-
quiring care in comparing records. Nonetheless,
average global curves for some elements have
been compiled and appear to work quite well
(e.g., Halverson et al., 2005; Saltzman, 2005;
Zhu et al., 2006). Some stratigraphic positions,
such as the base of the Cambrian (Corsetti and
Hagadorn, 2000; Amthor et al., 2003; Bowring
et al., 2007; Babcock et al., 2011), and the base
of the Oligocene (Zachos et al., 2001), as well as
some geologic events such as the end-Permian
extinction (e.g., Shen et al., 2011; Cao et al.,
2009) are marked by sharp, short-lived pertur-
bations in seawater chemistry that can be used
to correlate disparate sections.
Light stable isotopes and, in particular, varia-
tions in 13C/12C and 18O/16O are the most widely
applied chemostratigraphic tools, largely be-
cause of their relative low acquisition costs. The
smooth trends in their global values over long
periods of time, punctuated by large, mostly
short-lived and globally synchronous perturba-
tions, aid in correlation. Also used is the heavier
system of 87Sr/86Sr, although data acquisition
is more time consuming and expensive. Rock
magnetic stratigraphy may utilize several rock
magnetic parameters, including but not limited
to bulk magnetic susceptibility, intensity of an-
hysteretic remanent magnetization (ARM), in-
tensity of isothermal remanent magnetization
(IRM), including saturation IRM (SIRM), and
S-ratios (IRM acquired by fi rst saturating the
sample in a high fi eld and then applying a back-
fi eld IRM in a fi eld of –300 mT; this is divided
by the SIRM).
Isotopes of Carbon
It is generally assumed that the carbon iso-
topic composition of carbonate rocks (expressed
as δ13Ccarb) is set during equilibrium precipita-
tion from seawater and refl ects the isotopic
composition of the contemporaneous dissolved
inorganic carbon reservoir, which is in turn set
by the relative fraction of inorganic to organic
carbon burial (e.g., Kump and Arthur, 1999).
Excursions from δ13Ccarb = 0 (e.g., Shields and
Veizer, 2002; Maloof et al., 2010; Zachos et al.,
2001) can be used for global correlation, espe-
cially when calibrated with high-precision geo-
chronology. For the Neoproterozoic, the large
amplitude of well-documented global δ13Ccarb
excursions has led to controversy over the
canon ical interpretation of these records. Like
the Phanerozoic, positive excursions are thought
to refl ect enhanced organic carbon burial. The
cause of the negative excursions is debated, and
hypotheses include reconstitution of organic
carbon (e.g., Rothman et al., 2003; Fike et al.,
2006) and catastrophic methane release (e.g.,
Kennedy et al., 2001).
Isotopes of Oxygen
Secular variations in marine 16O/18O, denoted
as δ18O, have been a very important chemo-
stratigraphic tool, especially in paleoclimate
studies over the last 55 Ma (e.g., Zachos et al.,
2001). The basis for the method is that 16O is
preferentially enriched in water vapor leaving
the oceans, resulting in seawater that is rela-
tively enriched in 18O. Organisms that precipi-
tate shells as well as CaCO3 refl ect the oxygen
isotope composition of seawater. When the con-
tinents have ice sheets, the ice is enriched in 16O
from precipitation, and the oceans are depleted
in 16O, so that the δ18O is increased. During
times of no ice, precipitation is mixed back into
the oceans, and there is little change in δ18O.
Temperature also controls the δ18O of precipita-
tion, with higher fractionations correlated with
lower mean annual temperatures. It has long
been appreciated that the oxygen isotope record
shows cyclic variation, and Emiliani (1955) was
the fi rst to argue that these variations over the
past 55 Ma have been the result of oscillations
between glacial and interglacial stages. Thus,
the oxygen isotopic composition of minerals
precipitated from seawater can be used to infer
ice volumes and ocean temperatures.
Variations in the oxygen isotopic composi-
tion of foraminifera have played an important
role in the calibration of the astronomical time
scale, related to the periods of orbital eccen-
tricity, obliquity, and precession (see detailed
discussion in the following section). Tilt and
precession control the seasonal variations in
solar radiation. This has led to dramatic varia-
tions in the temperature of the Northern Hemi-
sphere over the last 25 ka. Using the variations
in oxygen isotopes of calcite in foraminiferan
shells, it is possible to reconstruct past ice vol-
umes and temperatures. Ice volume depends
largely on the amount of summer radiation at
high latitudes in the Northern Hemisphere. Most
of the variation in δ18O can be related to orbital
forcing of summer radiation. During glaciations,
there is less seasonal variability, the Earth-Sun
distance is high, and Earth’s orbit is character-
ized by high eccentricity and low tilt. During
interglacial periods , there is strong seasonality,
the Earth-Sun distance is lower, and Earth’s orbit
is characterized by low eccentricity and high tilt.
These orbital variations have led to the recogni-
tion of ~50 different episodes of growth of con-
tinental ice or ice ages over the past 3 Ma and a
way to calibrate the astronomical time scale.
Oxygen isotopic studies of older rocks are
also of value and have been used with great
success in the Cretaceous. Unusually fresh
forami nifera from the Demerara Rise in the
western tropical Atlantic record the warmest
(~36 °C) sea-surface temperatures (SSTs) re-
ported for Cretaceous–Cenozoic time (Wilson
et al., 2002). This in turn supports the idea of a
Cretaceous greenhouse. Other studies that have
reconstructed Cretaceous paleotemperatures us-
ing oxygen isotopes include Norris and Wilson
(1998), Bice et al. (2003), and Huber et al.
(2002). The further back in time one goes using
oxygen isotopes, the more carefully diagenetic
effects must be evaluated.
Isotopes of Strontium
Over much of Earth history, the 87Sr/86Sr
ratio of seawater has varied (e.g., Veizer, 1989;
Halver son et al., 2007), and it can be used to
evaluate the relative contributions of the major
sources of dissolved Sr, namely, riverine input
from chemical weathering of (1) continental
rocks that range from old radiogenic Sr from
shield areas and active orogens to (2) relatively
on August 30, 2015gsabulletin.gsapubs.orgDownloaded from
Walker et al.
268 Geological Society of America Bulletin, March/April 2013
unradiogenic Sr contained in weathered lime-
stones, and (3) the hydrothermal Sr fl ux, which
is similar to the MORB source. Samples amena-
ble to 87Sr/86Sr analysis include calcite, dolomite,
shells, and authigenic or biogenic phosphates,
especially conodonts. Diagenesis and contami-
nation by Sr residing in clay mineral lattices can
perturb the primary seawater signal of 87Sr/86Sr.
Calcite samples must be carefully screened for
high Sr concentrations, 100 ppm to >2000 ppm,
and in some cases, trace ele ments (e.g., Brand
and Veizer, 1980; Jacobsen and Kaufman, 1999).
Veizer (1989) provided a detailed review of
the origins of strontium in seawater and the
processes that affected the preservation of
the signal, including the effects of diagenesis.
He constructed the fi rst Phanerozoic Sr sea water
curve using CaCO3 from rocks and shells of dif-
ferent ages. He argued that the rate of mixing
of ocean waters and the long-residence time of
strontium in seawater could potentially yield a
reliable signal. He assumed that by using the
least radiogenic (i.e., lowest 87Sr/86Sr) ratios for
a given time and using large amounts of data,
one could minimize the effects of diagenesis
and produce a robust estimate of the isotopic
composition of seawater with time. McArthur
(1994, 1998) and McArthur and Howarth
(2004) have taken the lead in the development
and appli ca tion of the approach to age determi-
nation and standardization of the method, espe-
cially when there are large and unidirectional
changes, and its integration into the process of
refi ning the time scale.
The 87Sr/86Sr ratio of marine limestones can
be used for global correlations as well as inter-
pretations concerning the role of tectonics and
climate in determining the isotopic composi-
tion of seawater. For example, the dramatic rise
in 87Sr/86Sr after ca. 40 Ma (Burke et al., 1982;
Palmer and Elderfi eld, 1985; Edmond, 1992)
is attributed to the dissolved load (radiogenic
and high concentrations) derived from the ris-
ing and rapidly eroding Himalayas and carried
to the oceans by the Ganges and Brahmaputra-
Tsangpo River systems. Raymo and Ruddiman
(1992) suggested the rapid rise was related to
an enhanced rate of continental weathering and
thus climate. The Himalayas contain a broad
range of rocks with radiogenic Sr, especially
metamorphic rocks, which were likely the domi-
nant control on the Sr isotopic composition of
seawater at that time (Harris, 1995). It remains
unclear whether this trend can be used in deep
time to fi ngerprint the rise of ancient mountain
ranges or to make inferences about climate.
However, for rocks with good radioisotopic
and biostratigraphic control, such a rise can be
used as a chronometer with precisions of ~1 Ma
or better. Very detailed and precise strontium
isotope records exist for the Cenozoic (e.g.,
Paytan et al., 1993) and Mesozoic (McArthur
et al., 2001), but the resolution of the record
decreases further back in Earth history (Veizer
et al., 1999; Shields and Veizer, 2002; Halverson
et al., 2007).
Chemostratigraphy in the Precambrian
The most recent addition of a new geologic
period occurred in 2004 (Knoll et al., 2004)
when the Ediacaran Period was defi ned with a
GSSP in southern Australia. This period brack-
ets the period of time from one of the last major
ice ages (Marinoan) to the base of the Cambrian.
Because of the dearth of fossils with well-known
ranges in this period and the global nature of the
bounding events, carbon isotope chemostratig-
raphy (Ediacaran-Cambrian transition) and the
signal of deglaciation (Cryogenian-Ediacaran
boundary) were used to defi ne the newest period
(Knoll et al., 2004, 2006). These global isotope
excursions, interpreted to be isochronous, and
events such as the global glaciations that char-
acterize the Neoproterozoic offer much hope for
further subdivision of Earth history.
In summary, several chemostratigraphic
methods provide very powerful tools for global
correlation between radioisotopic tie points.
Moreover, in addition to providing a framework
for relative time, chemostratigraphic variations
are ultimately driven by evolving seawater com-
position, and thus, when integrated with the fos-
sil and tectonic records, great insight may be
gained in understanding the history of oceanic
circulation and timing of major changes in
ocean chemistry.
Astrochronology
There are significant gravitational inter-
actions between Earth and the Sun, Moon, and
other planets. Earth’s orientation relative to the
Sun changes in a quasi-periodic fashion as a re-
sult of interactions between Earth’s axial preces-
sion and the variable shape of its orbit induced
by motions of other planets (Berger and Loutre,
1990; Laskar et al., 2004; Hinnov and Hilgen,
2012). The interactions result in cyclic oscilla-
tions in the eccentricity of Earth’s orbit and in
the tilt and precession of Earth’s axis. Orbital
eccentricity variations have periods of ~100,000
years and 405,000 years, whereas the tilt varia-
tion has a period of 41,000 years, and the “cli-
matic precession” has a period of ~21,000 years,
and apart from the 405,000-year cycles, there
are multiple periods for each of these modes.
These in turn cause variations in solar radia-
tion reaching Earth’s surface, which result in
climatic variations, called Milankovitch cycles.
The climatic variations are recorded as cycli-
cally deposited sediments, which can be tuned
to the calculated Milankovitch cycles together
with stratigraphic correlation tools, principally
biostratigraphy and magnetostratigraphy.
This has developed into the use of “target
curves,” which provide numerical age calibra-
tion, with no reliance on radioisotopic geo-
chronol ogy, by matching local cyclostratigraphic
records with a highly constrained model of the
astronomical parameters or the climatic expres-
sion of those parameters such as variations in
solar insolation (the target curve). In the past
decade , dramatic improvements in understand-
ing solar system dynamics have resulted in the
extension of sophisticated models that predict
the climatic and sedimentologic response of
Earth to the astronomical parameters at least as
far back as 40–50 Ma (Laskar et al., 2011).
A similar approach for older rocks that can-
not be directly referenced to the late Cenozoic
allows the use of “fl oating” segments of cycli-
cally deposited sedimentary rocks to estimate
the amount of time between two stratigraphic
horizons. There are many examples of spec-
tacular cycles preserved in both marine and
continental sedimentary rocks that are defi ned
by, for example, color, geochemistry, rock mag-
netic properties, sedimentary structures, and
organic carbon content. The signal is measured
and assigned relative ages using magnetic po-
larity stratigraphy and biostratigraphy, and
then, if possible, calibrated in numerical time
with high-precision geochronologic informa-
tion on volcanic ash beds. These sections can
be used to correlate with other sections that
may not have cyclostratigraphy but include the
same biostratigraphy or magnetic polarity stra-
tigraphy. When combined with geochronology
of volcanic horizons, one can use the cycles
to interpolate time with higher precision than
with geochronology alone or linear interpola-
tion assuming constant sediment accumulation
rates. This has been used for sequences in the
Cretaceous (e.g., Meyers et al., 2012) and in
much older strata, such as those deposited near
the Triassic-Jurassic transition (e.g., Whiteside
et al., 2007) and in the Carboniferous (Davydov
et al., 2010).
The target curve and fl oating curve calibration
approaches have been important for developing
an astronomical time scale in the construction of
the global geologic time scale back to 250 Ma
(Gradstein et al., 2012). The astronomical time
scale alone has a resolution of 0.02–0.40 m.y.
for cyclostratigraphy in Neogene and Quater-
nary sections, and at the low end, it is com-
para ble to radioisotopic dating but completely
independent. For older rocks, the astronomical
time scale can provide age models between
radio iso topi cally dated tie points allowing fi ne-
on August 30, 2015gsabulletin.gsapubs.orgDownloaded from
The Geological Society of America Geologic Time Scale
Geological Society of America Bulletin, March/April 2013 269
scale resolution in thick sequences lacking other
geochronologic proxies.
Parallel improvements in both 40Ar/39Ar and
U-Pb geochronology have made it possible to
pursue the intercalibration of radioisotopic dat-
ing and astrochronology (e.g., Kuiper et al.,
2008; Meyers et al., 2012). When successful,
this integration allows for unprecedented ac-
curacy and precision in the measurement of
geologic time. The basic approach is to use
high-precision geochronology of volcanic ash
beds with both U-Pb and 40Ar/39Ar techniques
to yield an estimate of the duration between
each dated deposit. Despite the complexities
involving systematic differences between U-Pb
and 40Ar/39Ar dates, the age difference between
each deposit should be the same but likely with
different uncertainties. Then, the fl oating astro-
nomical model for the time interval can be
used for a much higher-resolution age model.
Ideally , bedded sedimentary rock sequences
with orbitally infl uenced cycles occur nearby or
between the dated ash beds, so that one can be
sure the fl oating time scale can be applied to a
particular interval. This is a rare occurrence but
allows an opportunity to directly intercalibrate
two independent radioisotopic chronometers
against an astrochronologic age model (Meyers
et al., 2012).
Another important application was reported
by Kuiper et al. (2008), who used 40Ar/39Ar
geochronology to determine the age of the Fish
Canyon sanidine by astronomically calibrating
high-precision dates on ash beds and arriving
at a value of 28.201 ± 0.046 Ma. Meyers et al.
(2012) directly tested this by comparing U-Pb
and 40Ar/39Ar dates from the same bentonites
and concluded that the pairs of dates, using
an assumed age for Fish Canyon sanidine of
28.201 Ma, were the same within uncertainty.
They inte grated the 40Ar/39Ar dates from three
bentonites and the orbital time scale with un-
certainties of thousands of years to confi rm that
28.201 Ma is the best age estimate for Fish Can-
yon sanidine. Renne et al. (2010, 2011) arrived
at a slightly older age estimate (28.291 Ma ±
0.036) by comparing pairs of U-Pb zircon and
40Ar/39Ar dates from rocks from a wide range of
ages with the Fish Canyon sanidine age proposed
by Kuiper et al. (2008). In future intercalibration
studies, the sources of uncertainty associated
with the radioisotopic ages (analytical, tracer so-
lution, or age of neutron fl uence standard, and
decay constants), as well as uncertainties in the
astro chronol ogy, must be considered.
A Geologic Time Scale 2012 (Gradstein et al.,
2012) uses “absolute” astronomical calibrations
based on solar system dynamics for the Neo-
gene, and “fl oating” astronomical calibrations
tied to radioisotopically dated stratigraphic
points for the Paleogene and all three Meso-
zoic periods. The geologic time scale includes
absolute astronomical calibrations for the entire
Cenozoic Era and Cretaceous Period; fl oating
astronomical calibrations make up 85% of the
Jurassic and 75% of the Triassic time scales.
The eccentricity of Earth’s orbit has a major
variation with a period of 405,000 years.
Shorter astronomical cycles with precession and
obliquity periods are used to calibrate Neogene
and younger rocks. However, because the solar
system exhibits chaotic diffusion, we can expect
that their periods will change with time beyond
50 m.y., and the short cycles cannot be used to
calibrate the older (>50–60 Ma) rock record. The
405,000-year eccentricity cycle remains more
stable through geologic time because it results
from interactions of Jupiter and Venus. Workers
are now using fl oating cyclostratigraphic rec-
ords that contain what they interpret to be the
405,000-year cycle to calibrate the geologic time
scale dating back to at least the end of the Paleo-
zoic era (ca. 252.2 Ma) and beyond.
There is potential to use the fl oating curve
approach to greatly increase the precision of
the geologic time scale. For example, Davydov
et al. (2010) used U-Pb geochronology in the
Donets Basin to confi rm that individual high-
frequency Carboniferous cyclothems and bun-
dles of cyclothems into fourth-order sequences
are the eustatic response to orbital eccentricity
(100 and 405 ka) forcing. They produced a con-
tinuous age model for Pennsylvanian strata of
the basin during this key interval of Earth his-
tory characterized by a major ice age. Detailed
biostratigraphy will allow the export of this age
model to sections throughout Euramerica.
This approach has also been used in the re-
markable sedimentary sequence of dominantly
lacustrine facies of the Newark Basin, where in-
vestigation of the continuous sequence obtained
by a series of drill cores has allowed unprece-
dented resolution of a fl oating astronomical time
scale for over 5800 m of composite stratigraphy
integrated with magnetic polarity stratig raphy
that includes 59 magnetozones and a sampling
density of about every 20 ka (Kent and Olsen,
1999; Olsen et al., 2011). Vertebrate biostratig-
raphy and geochronology of the section are
sparse, although several Central Atlantic mag-
matic province intrusions and lava fl ows at ca.
201.4 Ma, near the end-Triassic extinction, are
used as an upper tie point. There is a pronounced
405 ka cyclicity, and it has been used for specu-
lations on the tempo and mode of dinosaurian
origin, diversifi cation, and rise to ecological
dominance (Olsen et al., 2011).
A good example of combining magnetic
polarity information and geochronology and
testing the results against astrochronological
models was recently published by Tsukui and
Clyde (2012). In this paper, a highly resolved,
previously obtained geomagnetic polarity time
scale for the early-middle Eocene was combined
with 40Ar/39Ar sanidine age determinations on
23 discrete ash-fall tuffs and other chronostrati-
graphic information, including polarity data, to
support a revised paleomagnetic time scale for
the early to middle Eocene. The age calibration
model suggests an approximate 2 Ma duration
for the early Eocene climatic optimum, which
in turn coincides with the inferred age of the
Wasatchian-Bridgerian faunal transition. Com-
pared with a previous astrochronological model,
the revised time scale is consistent with an age
for the Paleocene-Eocene thermal maximum of
56.33 Ma. Ultimately, terrestrial sequences with
relatively high accumulation rates and high-
precision geochronology can better resolve the
age and duration of polarity chrons, and can
be compared with astrochronologic data from
more continuous marine sedimentary deposits.
In summary, the integration of high-preci-
sion, high-accuracy geochronology with refi ned
astrochronological models and stable isotope,
chemostratigraphic, and magnetostratigraphic
observations will allow us to better understand
the role of climate and ocean-atmosphere inter-
actions on biological evolution.
OUTSTANDING ISSUES
Stratigraphic Challenges
The search for satisfactory horizons to use for
marking GSSPs in the Phanerozoic has been a
long protracted process, and many refi nements
in strategy and technical methods have been
introduced since the time that the fi rst GSSP,
which marked the base of the Devonian, was
ratifi ed in 1972. With the appearance of new
information (particularly detailed chemostrati-
graphic data) and reevaluation of sections, some
early decisions as to specifi c time boundaries
now seem less well justifi ed, and may merit
reexamination. One example is the basal Cam-
brian GSSP. Because of problems associated
with global correlation of the Cambrian/Paleo-
zoic-Phanerozoic base using the biostratigraphy
of trace fossils, chemostratigraphy has emerged
as the de facto means of characterizing and cor-
relating that horizon globally (Bowring et al.,
2007; Babcock et al., 2011). The Ediacaran-
Cambrian boundary in Newfoundland is based
on trace fossil assemblages and has no body
fossils and has no rocks suitable for geochronol-
ogy or chemostratigraphy. It remains uncertain
whether the negative δ13C excursion used to ap-
proximate the boundary does in fact coincide
with the Cambrian GSSP.
on August 30, 2015gsabulletin.gsapubs.orgDownloaded from
Walker et al.
270 Geological Society of America Bulletin, March/April 2013
Up to the present time, boundary positions in
the Archean and Proterozoic have mostly used
purely chronometric boundaries called global
standard stratigraphic ages (GSSAs), based on
geochronologic data, rather than points in sec-
tions or GSSPs. The one exception is the ter-
minal system/period of the Proterozoic, the
Ediacaran. Ultimately, it will be desirable, if pos-
sible, to replace GSSAs with GSSPs that have
well-calibrated numerical ages.
The recent defi nition of the base of the Holo-
cene Series using signals of climatic change
(Walker et al., 2009), rather than biostrati-
graphic markers, represents an important turn-
ing point in philosophy about GSSP defi nition.
Some nonbiostratigraphic proxies are superior
to biostratigraphic tools in certain parts of the
geologic column, and it is likely that nonbio-
stratigraphic correlation techniques will play an
increased role in the defi nition of GSSPs in the
future, particularly those that will be defi ned in
the pre-Phanerozoic.
Geochronology and Resolving
Discrepancies between U-Pb and
40Ar/39Ar Methods
Ultimately, the geoscience community will
succeed in building a time scale that can seam-
lessly use both high-precision 40Ar/39Ar data
and U-Pb geochronologic data. The concern
that we presently face, however, is one that has
been recognized for some time, and that is the
systematic bias between U-Pb and 40Ar/39Ar
dates, which in some cases approaches ~1%,
with U-Pb dates being consistently older. The
causes of this systematic bias are not clear
and likely include inaccuracies in the decay
constants for both systems, inaccuracies in
the age of fl uence monitors used in 40Ar/39Ar
geo chronol ogy, an inaccurate assumed value
for the 238U/235U ratio in zircons (Hiess et al.,
2012), and the possibility that in some vol-
canic rocks the U-Pb system in zircon does not
record the time of emplacement but rather a
short period of pre-eruptive residence. Because
all 40Ar/39Ar approaches are relative dating
methods, the age of the fl uence monitor must
be independently determined, and sanidine
from the Fish Canyon tuff is the most com-
monly used. Renne et al. (2010, 2011) inverted
data sets of U-Pb zircon and 40Ar/39Ar dates to
arrive at a more accurate estimate of the age
of Fish Canyon sanidine. This is a promising
approach as we acquire many additional high-
precision 40Ar/39Ar and U-Pb data sets on the
same samples from several different localities
representing more and more of geologic time.
The estimate of the assumed age of the Fish
Canyon sanidine has varied from ca. 27.7 Ma
to 28.29 Ma (Renne et al., 2010, 2011). Many
workers in the community are using 28.201 Ma
as the preferred age, as determined by astro-
nomical calibration (Kuiper et al., 2008),
as in many cases it appears to result in good
agreement between 40Ar/39Ar and U-Pb deter-
minations. The community is making great
progress, but more work remains to be done.
Our goal was to provide a synopsis of the
history of the Geological Society of America
Geologic Time Scale and the many different
facets of deep time investigations that have led
to refi nements in geologic time scales. This
overview highlights the continued importance
of integrated efforts to better understand the
Earth-life record of deep time on a global scale.
A crucial aspect of progress in geochronology
is the use of highly characterized mineral stan-
dards that span a range of ages. EARTHTIME
is actively seeking out standard samples that
range in age from less than 1 Ma to well over
500 Ma that can be analyzed for both 40Ar/39Ar
and U-Pb (Schoene et al., 2005; Renne et al.,
2010). Mineral separates will be distributed to
participating laboratories, and, after analyses
are completed, the results will be published.
Ideally, when a U-Pb laboratory publishes a
study, they will also publish data for one or
more standards run during the same interval as
the study, as is currently done in many 40Ar/39Ar
laboratories.
The Geological Society of America Geologic
Time Scale and Its Future
Palmer’s seminal work on the Geological So-
ciety of America Geologic Time Scale (Palmer,
1983) was important for the development of time
scales in general. This model has been adopted
by the ICS in the presentation of their geologic
time scale (International Commission on Stra-
tigraphy, 2012). The 1983 Geological Society of
America Geologic Time Scale was constructed
with an emphasis on North American geology.
The Geological Society of America will con-
tinue to follow the ICS and other efforts such as
Gradstein et al. (2012) in updating the chrono-
strati graphic scale and integrating it with the
chronometric one, and the GSA will maintain
its presentation of the time scale as a service to
its membership, the larger scientifi c community,
and the public to promote the scientifi c and edu-
cational goals of the society.
ACKNOWLEDGMENTS
We thank A.R. (Pete) Palmer for extensive dis-
cussion and encouragement. James Ogg shared pre-
prints of parts of the Gradstein et al. (2012) book.
J.C. Creveling , Linda Hinnov, Brad Cramer, and Paul
Olsen are thanked for comments.
REFERENCES CITED
Alvarez, L.W., Alvarez, W., Asaro, F., and Michel, H.V.,
1980, Extraterrestrial cause for the Cretaceous-Tertiary
extinction: Science, v. 208, p. 1095–1108, doi:10.1126
/science.208.4448.1095.
Amthor, J.E., Grotzinger, J.P., Schröder, S., Bowring, S.A.,
Ramezani, J., Martin, M.W., and Matter, A., 2003, Ex-
tinction of Cloudina and Namacalathus at the Precam-
brian-Cambrian boundary in Oman: Geology, v. 31,
p. 431–434, doi:10.1130/0091-7613(2003)031<0431:
EOCANA>2.0.CO;2.
Babcock, L.E., Robison, R.A., and Peng, S.C., 2011, Cam-
brian stage and series nomenclature of Laurentia and
the developing global chronostratigraphic scale: Mu-
seum of Northern Arizona Bulletin, v. 67, p. 12–26.
Barrell, J., 1917, Rhythms and the measurement of geologic
time: Geological Society of America Bulletin, v. 28,
p. 748–749.
Berger, A., and Loutre, M.F., 1990, Origine des fréquences
des eléments astronomiques intervenant dans le calcul
de l’insolation: Bulletin de la Classe des Sciences,
Académie Royale Belgique, Ser. 6 1(1[3]), p. 45–106.
Berggren, W.A., Kent, D.V., Flynn, J.J., and Van Couver-
ing, J.A., 1985a, Cenozoic geochronology: Geologi-
cal Society of America Bulletin, v. 96, p. 1407–1418,
doi:10.1130/0016-7606(1985)96<1407:CG>2.0.CO;2.
Berggren, W.A., Kent, D.V., and Van Couvering, J.A., 1985b,
The Neogene: Part 2. Geochronology and chrono-
stratigraphy, in Snelling, N.J., ed., The Chronology of
the Geological Record: Geological Society of London
Memoir 10, p. 211–260.
Bice, K.L., Huber, B.T., and Norris, R.D., 2003, Extreme
polar warmth during the Cretaceous greenhouse? Para-
dox of the late Turonian δ18O record at Deep Sea Drill-
ing Project Site 511: Paleoceanography, v. 18, 1031,
doi:10.1029/2002PA000848.
Bowring, J.F., McLean, N.M., and Bowring, S.A., 2011,
Engineering cyber infrastructure for U-Pb geochro-
nology: Tripoli and U-Pb_Redux: Geochemistry Geo-
physics Geosystems, v. 12, Q0AA19, doi:10.1029
/2010GC003479.
Bowring, S.A., Grotzinger, J.P., Condon, D.J., Ramezani, J.,
Newall, M.J., and Allan, P.A., 2007, Geochronologic
constraints on the chronostratigraphic framework of the
Neoproterozoic HUQF Supergroup, Sultanate of Oman:
American Journal of Science, v. 307, p. 1097–1145
http://dx.DOI.org/10.2475/10.2007.01.
Brand, U., and Veizer, J., 1980, Chemical diagenesis of a
multicomponent carbonate system: 1. Trace elements:
Journal of Sedimentary Petrology, v. 50, p. 1219–1236.
Brunhes, B., 1906, Recherches sur le direction d’aimenta-
tion des roches volcaniques: Journal of Physics, v. 5,
p. 705–724.
Burke, W., Denison, R., Hetherington, E., Koepnik, R.,
Nelson, M., and Omo, J., 1982, Variations of sea water
87Sr/86Sr throughout Phanerozoic shales: Geology,
v. 10, p. 516–519, doi:10.1130/0091-7613(1982)10
<516:VOSSTP>2.0.CO;2.
Cande, S.C., and Kent, D.V., 1992, A new geomagnetic
polarity time scale for the Late Cretaceous and
Ceno zoic: Journal of Geophysical Research, v. 97,
p. 13,917-13,951.
Cande, S.C., and Kent, D.V., 1995, Revised calibration of
the geomagnetic polarity timescale for the Late Creta-
ceous and Cenozoic: Journal of Geophysical Research,
v. 100, p. 6093–6095, doi:10.1029/94JB03098.
Cao, C. Love, G.D., Hays, L.E., Wanga, W., Shen, S., and
Summons, R.E., 2009, Biogeochemical evidence for
euxinic oceans and ecological disturbance presaging
the end-Permian mass extinction event: Earth and Plan-
etary Science Letters, v. 281, p. 188–201, doi:10.1016
/j.epsl.2009.02.012.
Cohen, K.M., Finney, S., and Gibbard, P.L., 2012, Inter-
national Chronostratigraphic Chart: International
Commission on Stratigraphy, www.stratigraphy.org
(last accessed May 2012). (Chart reproduced for the
34th International Geological Congress, Brisbane,
Australia, 5–10 August 2012.)
Condon, D., Schoene, B., Bowring, S.A., Parrish, R.,
McLean, N., Noble, S., and Crowley, Q., 2007,
EARTHTIME: Isotopic tracers and optimized solutions
on August 30, 2015gsabulletin.gsapubs.orgDownloaded from
The Geological Society of America Geologic Time Scale
Geological Society of America Bulletin, March/April 2013 271
for high-precision U-Pb ID-TIMS geochronology: Eos
(Transactions, American Geophysical Union), v. 88,
no. 52, abs. V41E–06.
Corsetti, F.A., and Hagadorn, J.W., 2000, Precambrian-
Cambrian transition: Death Valley, United States:
Geology v. 28, p. 299–302, doi: 10.1130/0091-7613
(2000)28<299:PTDVUS>2.0.CO;2
Davydov, V.I., Crowley, J.L., Schmitz, M.D., and Poletaev,
V.I., 2010, High-precision U-Pb zircon age calibration
of the global Carboniferous time scale and Milanko-
vitch band cyclicity in the Donets Basin, eastern
Ukraine: Geochemistry Geophysics Geosystems, v. 11,
Q0AA04, doi:10.1029/2009GC002736.
Doell, R.R., and Dalrymple, G.B., 1966, Geomagnetic
polarity epochs: A new polarity event and the age of
the Brunhes Matuyama boundary: Science, v. 152,
p. 1060–1061, doi:10.1126/science.152.3725.1060.
Eardley, A.J., 1951, Structural Geology of North America:
New York, Harper and Brothers, 624 p.
Edmond, J.M., 1992, Himalayan tectonics, weathering
processes and the strontium isotope record in marine
limestones: Science, v. 258, p. 1594–1597, doi:10.1126
/science.258.5088.1594.
Emiliani, C., 1955, Pleistocene temperatures: The Journal of
Geology, v. 63, p. 538–578, doi:10.1086/626295.
Evans, M.E., and Heller, F., 2003, Environmental Magne-
tism: Principles and Applications of Enviromag-
netics: San Diego, California, Academic Press, 299 p.,
doi:10.1002/jqs.858.
Fike, D.A., Grotzinger, J.P., Pratt, L.M., and Summons, R.E.,
2006, Oxidation of the Ediacaran ocean: Nature, v. 444,
p. 744–747, doi:10.1038/nature05345.
Glen, W., 1982, The Road to Jaramillo: Stanford, California,
Stanford University Press, Stanford, 459 p.
Gradstein, F.M., and 38 others, 2004, A Geologic Time
Scale 2004: Cambridge, UK, Cambridge University
Press, 589 p.
Gradstein, F.M., Ogg, J.G., Schmitz, M.D., and Ogg, G.M.,
editors, 2012, The Geologic Time Scale 2012, vol. 1:
Boston, Elsevier, 1144 p., http://dx.doi.org/10.1016
/B978-0-444-59425-9.01001-5.
Halverson, G.P., Hoffman, P.F., Schrag, D.P., Maloof, A.C.,
and Rice, A.H., 2005, Towards a Neoproterozoic com-
posite carbon-isotope record: Geological Society of
America Bulletin, v. 117, p. 1181–1207, doi:10.1130
/B25630.1.
Halverson, G.P., Dudas, F.O., Maloof, A.C., and Bowring,
S.A., 2007, Evolution of the 87Sr/86Sr composition of
Neoproterozoic seawater: Palaeogeography, Palaeo-
climatology, Palaeoecology, v. 256, p. 103–129, doi:
10.1016/j.palaeo.2007.02.028.
Harland, W.B., Smith, A.G., and Wilcock, B., 1964, The
Phanerozoic Time-Scale—A Symposium Dedicated
to Professor Arthur Holmes: Quarterly Journal of the
Geological Society of London, v. 120S, 458 p.
Harland, W.B., Cox, A.V., Llewellyn, P.G., Picton, C.A.G.,
Smith, A.G., and Walters, R.W., 1982, A Geologic
Time Scale: Cambridge, UK, Cambridge University
Press, 131 p.
Harland, W.B., Cox, A.V., Llewellyn, P.G., Smith, A.G., and
Walters, R., 1990, A Geologic Time Scale 1989: Cam-
bridge, UK, Cambridge University Press.
Harris, N., 1995, Signifi cance of weathering Himalayan
metasedimentary rocks and leucogranites for the Sr
isotope evolution of seawater during the early Mio-
cene: Geology, v. 23, p. 795–798, doi:10.1130/0091
-7613(1995)023<0795:SOWHMR>2.3.CO;2.
Hiess, J., Condon, D.J., McLean, N., and Noble, S.R., 2012,
238U/235U systematics in terrestrial uranium-bearing
minerals: Science, v. 335, p. 1610–1614, doi:10.1126
/science.1215507.
Hilgen, F.J., 2010, Astronomical tuning in the 19th century:
Earth-Science Reviews, v. 98, p. 65–80
Hinnov, L.A., and Hilgen F.J., 2012, Cyclostratigraphy
and astrochronology, in Gradstein, F.M., Ogg, J.G.,
Schmitz, M.D., and Ogg, G.M., eds., The Geologic
Time Scale 2012: vol. 1: Boston, Elsevier, p. 63–83,
doi:10.1016/B978-0-444-59425-9.00004-4.
Holmes, A., 1913, The Age of the Earth: London and New
York, Harper, 196 p.
Holmes, A., 1937, The Age of the Earth (3rd ed.): London,
Nelson, 263 p.
Holmes, A., 1962, ‘Absolute age’ a meaningless term:
Nature , v. 196, p. 1238, doi:10.1038/1961238b0.
Huber, B.T., Norris, R.D., and MacLeod, K.G., 2002, Deep
sea paleotemperature record of extreme warmth during
the Cretaceous: Geology, v. 30, p. 123–126, doi:10.1130
/0091-7613(2002)030<0123:DSPROE>2.0.CO;2.
Hutton, J., 1788, Theory of the Earth; or an investigation of
the laws observable in the composition, dissolution, and
restoration of land upon the Globe: Transactions of the
Royal Society of Edinburgh, v. 1, Part 2, p. 209–304.
International Commission on Stratigraphy, 2012, Inter-
national Chronostratigraphic Chart: www.stratigraphy
.org (last accessed May 2012).
Irving, E., 1964, Paleomagnetism and its Applications to
Geological and Geophysical Problems: New York,
John Wiley and Sons, 399 p.
Irving, E., and Parry, L.G., 1963, The magnetism of some
Permian rocks from New South Wales: Geophysical
Journal of the Royal Astronomical Society, v. 7, p. 395–
411, doi:10.1111/j.1365-246X.1963.tb07084.x.
Jacobsen, S.B., and Kaufman, A.J., 1999, The Sr, C, and O
isotopic evolution of Neoproterozoic seawater: Chemi-
cal Geology, v. 161, p. 37–57, doi:10.1016/S0009-2541
(99)00080-7.
Kennedy, M.J., Christie-Blick, N., and Sohl, L.E., 2001,
Are Proterozoic cap carbonates and isotopic excur-
sions a record of gas hydrate destabilization following
Earth’s coldest intervals?: Geology, v. 29, p. 443–446,
doi:10.1130/0091-7613(2001)029<0443:APCCAI
>2.0.CO;2.
Kent, D.V., and Gradstein, F.M., 1986, A Jurassic to Recent
chronology, in Vogt, P.R., and Tucholke, B.E., eds.,
The Western North Atlantic Region: Boulder, Colo-
rado, Geological Society of America, The Geology of
North America, v. M, p. 45–50.
Kent, D.V., and Olsen, P.E., 1999, Astronomically tuned geo-
magnetic polarity time scale for the Late Triassic: Jour-
nal of Geophysical Research, v. 104, p. 12,831–12,841.
Kent, D.V., Olsen, P.E., and Witte, W.K., 1995, Late Trias-
sic–Early Jurassic geomagnetic polarity and paleolati-
tudes from drill cores in the Newark rift basin (eastern
North America): Journal of Geophysical Research,
v. 100, p. 14,965–14,998, doi:10.1029/95JB01054.
King, P.B., 1959, The Evolution of North America: Prince-
ton, New Jersey, Princeton University Press, 189 p.
Knoll, A.H., and Walter, M.R., 1992, Latest Proterozoic stra-
tigraphy and Earth history: Nature, v. 356, p. 673–678,
doi:10.1038/356673a0.
Knoll, A.H., Walter, M.R., Narbonne, G.M., and Christie-
Blick, N., 2004, A new period for the geologic time
scale: Science, v. 305, p. 621–622, doi:10.1126/science
.1098803.
Knoll, A.H., Walter, M.R., Narbonne, G.M., and Christie-
Blick, N., 2006, The Ediacaran Period: A new addition
to the geologic time scale: Lethaia, v. 39, p. 13–30,
doi:10.1080/00241160500409223.
Kodama, K.P., Anastasio, D.J., Newton, M.L., Pares, J.M.,
and Hinnov, L.A., 2010, High-resolution rock magnetic
cyclostratigraphy in an Eocene fl ysch, Spanish Pyre-
nees: Geochemistry Geophysics Geosystems, v. 11,
QOAA07, doi:10.1029/2011GC003069.
Korte, C., and Kozur, H., 2010, Carbon-isotope stratigraphy
across the Permian-Triassic boundary: A review: Journal
of Asian Earth Sciences, v. 39, p. 215–235, doi:10.1016
/j.jseaes.2010.01.005.
Kuiper, K.F., Deino, A., Hilgen, F.J., Krijgsman, W., Renne,
P.R., and Wijbrans, J.R., 2008, Synchronizing rock
clocks of Earth history: Science, v. 320, p. 500–504,
doi:10.1126/science.1154339.
Kulp, J.L., 1961, Geologic time scale: Science, v. 133,
p. 1105–1114, doi:10.1126/science.133.3459.1105.
Kump, L.R., and Arthur, M.A., 1999, Interpreting carbon-
isotope excursions: Carbonates and organic matter:
Chemical Geology, v. 161, p. 181–198, doi:10.1016
/S0009-2541(99)00086-8.
Laskar, J., Robutel, P., Joutel, F., Gastineau, M., Correia,
A.C.M., and Levrard, B., 2004, A long-term numeri-
cal solution for the insolation quantities of the Earth:
Astronomy & Astrophysics, v. 428, p. 261–285, doi:
10.1051/0004-6361:20041335.
Laskar, J., Fienga, A., Gastineau, M., and Manche, H., 2011,
La2010: A new orbital solution for the long term mo-
tion of the Earth: Astronomy & Astrophysics, v. 532,
doi:10.1051/0004-6361/201116836.
Li, H.M., and Wang, J.D., 1989, Magnetostratigraphy
of Permo-Triassic boundary section of Meishan of
Changxing: Zhejiang, Scientia Sinica, Series B, v. 32,
p. 1401–1408.
Lucas, S.G., Tanner, L.H., Donohoo-Hurley, L., Geissman,
J.W., Kozur, H.W., Heckert, A.B., and Weems, R.E.,
2011, Position of the Triassic-Jurassic boundary and
timing of the end-Triassic extinctions on land: Data
from the Moenave Formation on the southern Colo-
rado Plateau, USA: Palaeogeography, Palaeoclimatol-
ogy, Palaeoecology, v. 302, p. 194–205, doi:10.1016
/j.palaeo.2011.01.009.
Maloof, A.C., Porter, S.M., Moore, J.L., Dudás, F.Ö.,
Bowring , S.A., Higgins, J.A., Fike, D.A., and Eddy,
M.P. 2010, The earliest Cambrian record of animals
and ocean geochemical change: Geological Society of
America Bulletin, v. 122, p. 1731–1774, doi:10.1130
/B30346.1.
Mattinson, J.M., 2005, Zircon U-Pb chemical abrasion
(“CA-TIMS”) method: Combined annealing and multi-
step partial dissolution analysis for improved preci-
sion and accuracy of zircon ages: Chemical Geology,
v. 220, p. 47–66, doi:10.1016/j.chemgeo.2005.03.011.
Matuyama, M., 1929, On the direction of magnetization of
basalts in Japan, Tyosen, and Manchuria: Proceedings
of the Imperial Academy (Tokyo), v. 5, p. 203–205.
McArthur, J.M., 1994, Recent trends in strontium isotope
stratigraphy: Terra Nova, v. 6, p. 331–358, doi:10.1111
/j.1365-3121.1994.tb00507.x.
McArthur, J.M., 1998, Strontium isotope stratigraphy, in
Doyle, P., and Bennett, M.R., eds., Unlocking the
Stratigraphical Record: Chichester, UK, John Wiley
and Sons, p. 221–241.
McArthur, J.M., and Howarth, R.J., 2004, Strontium isotope
stratigraphy, in Gradstein, F., Ogg, J., and Smith, A.,
eds., A Geological Time Scale 2004: Cambridge, UK,
Cambridge University Press, p. 96–105.
McArthur, J.M., Howarth, R.J., and Bailey, T.R., 2001,
Strontium isotope stratigraphy: LOWESS Version 3:
Best fi t to the marine Sr-isotope curve for 0–509 Ma
and accompanying look-up table for deriving numeri-
cal age: The Journal of Geology, v. 109, p. 155–170,
doi:10.1086/319243.
McLean, N.M., Bowring, J.F., and Bowring, S.A., 2011,
An algorithm for U-Pb isotope dilution data reduc-
tion and uncertainty propagation: Geochemistry Geo-
physics Geosystems, v. 12, Q0AA18, doi:10.1029
/2010GC003478.
Meyers, S., Siewert, S.E., Singer, B.S., Sageman, B.B.,
Condon, D.J., Obradovich, J.D., Jicha, B.R., and
Sawyer, D.A., 2012, Intercalibration of radioisotopic
and astrochronologic time scales for the Cenomanian-
Turonian boundary interval, Western Interior Basin,
USA: Geology, v. 40, p. 7–10, doi:10.1130/G32261.1.
Norris, R.D., and Wilson, P.A., 1998, Low-latitude sea-
surface temperatures for the mid-Cretaceous and the
evolution of planktonic foraminifera: Geology, v. 26,
p. 823–826, doi:10.1130/0091-7613(1998)026<0823:
LLSSTF>2.3.CO;2.
Odin, G.S., 1982, Numerical Dating in Stratigraphy:
Chichester , UK, John Wiley, 1094 p.
Ogg, J.G., Ogg, G., and Gradstein, F.M., 2008, The Con-
cise Geologic Time Scale: Cambridge, UK, Cambridge
University Press, 177 p.
Olsen, P.E., Kent, D.V., Cornet, B., Witte, W.K., and Schlische,
R.W., 1996a, High-resolution stratigraphy of the Newark
Rift Basin (early Mesozoic, eastern North America):
Geological Society of America Bulletin, v. 108,
p. 40–77, doi:10.1130/0016-7606(1996)108<0040:
HRSOTN>2.3.CO;2.
Olsen, P.E., Schlische, R.W., and Fedosh, M.S., 1996b,
580 ky duration of the Early Jurassic fl ood basalt event
in eastern North America estimated using Milankovitch
cyclostratigraphy, in Morales, M., ed., The Continen-
tal Jurassic: Museum of Northern Arizona Bulletin 60,
p. 11–20.
Olsen, P.E., Kent, D.V., and Whiteside, J.H., 2011, Implica-
tions of the Newark Supergroup-based astrochronology
and geomagnetic polarity timescale (Newark-APTS)
for the tempo and mode of the early diversifi cation
on August 30, 2015gsabulletin.gsapubs.orgDownloaded from
Walker et al.
272 Geological Society of America Bulletin, March/April 2013
of the Dinosauria: Earth and Environmental Science
Transactions of the Royal Society of Edinburg, v. 101,
p. 1–33, doi:10.1017/S1755691011020032.
Opdyke, N., and Channell, J.E.T., 1996, Magnetic Stratigra-
phy: San Diego, Academic Press, 346 p.
Opdyke, N., Roberts, J., Claoue-Long, J., Irving, E., and
Jones, P., 2000, Base of the Kiaman: Its defi nition and
global stratigraphic signifi cance: Geological Society of
America Bulletin, v. 112, p. 1315–1341, doi:10.1130
/0016-7606(2000)112<1315:BOTKID>2.0.CO;2.
Oreskes, N., ed., 2001, Plate Tectonics: An Insider’s History
of the Modern Theory of the Earth: Boulder, Colorado,
Westview Press, 424 p.
Palmer, A.R., 1983, The Decade of North American Geol-
ogy 1983 Geologic Time Scale: Geology, v. 11, p. 503–
504, doi:10.1130/0091-7613(1983)11<503:TDONAG
>2.0.CO;2.
Palmer, A.R., and Geissman, J.W., 1999, 1999 Geologic
Time Scale: Boulder, Colorado, Geological Society of
America, 1 p.
Palmer, M.R., and Elderfi eld, H., 1985, Strontium isotope
composition of sea water over the past 75 Myr: Nature,
v. 314, p. 526–528, doi:10.1038/314526a0.
Pavlov, V.E., and Gallet, Y., 2005, A third superchron during
the Early Paleozoic: Episodes, v. 28, no. 2, p. 1–7.
Pavlov,V.E., Veselovskiy, R.V., Shatsillo, A.V., and Gallet ,
Y., 2012, Magnetostratigraphy of the Ordovician
Angara/Rozhkova River Section: Further Evidence for
the Moyero Reversed Superchron: Izvestiya, Physics
of the Solid Earth, v. 48, p. 297–305.
Paytan, A., Kastner, M., Martin, E.E., Macdougall, J.D., and
Herbert, T., 1993, Marine barite as a monitor of sea-
water strontium isotope composition: Nature, v. 366,
p. 445–449, doi:10.1038/366445a0.
Raymo, M.E., and Ruddiman, W.F., 1992, Tectonic forcing
of late Cenozoic climate: Nature, v. 359, p. 117–122,
doi:10.1038/359117a0.
Renne, P.R., Mundil, R., Balco, G., Min, K., and Ludwig,
K.R., 2010, Joint determination of 40K decay constants
and 40Ar*/40K for the Fish Canyon sanidine standard,
and improved accuracy for 40Ar/39Ar geochronology:
Geochimica et Cosmochimica Acta, v. 74, no. 18,
p. 5349–5367, doi:10.1016/j.gca.2010.06.017.
Renne, P.R., Balco, G., Ludwig, K., Mundil, R., and Min,
K., 2011, Response to the comment by W.H. Schwarz
et al. on “Joint determination of 40K decay constants
and40Ar*/40K for Fish Canyon sanidine standard, and
improved accuracy for 40Ar/39Ar geochronology”
by P.R. Renne et al. (2010): Geochimica et Cosmo-
chimica Acta, v. 75, p. 5097–5100, doi:10.1016/j.gca
.2011.06.021.
Rothman, D.H., Hayes, J.M., and Summons, R.E., 2003,
Dynamics of the Neoproterozoic carbon cycle: Pro-
ceedings of the National Academy of Sciences of
the United States of America, v. 100, p. 124–129,
doi:10.1073/pnas.0832439100.
Saltzman, M.R., 2005, Phosphorus, nitrogen, and the redox
evolution of the Paleozoic oceans: Geology, v. 33,
p. 573–576, doi:10.1130/G21535.1.
Salvador, A., ed., 1994, International Stratigraphic Guide
(2nd ed.): Trondheim, Norway, and Boulder, Colorado,
USA, International Union of Geological Sciences and
the Geological Society of America, 214 p.
Schoene, B., Crowley, J.L., Condon, J.D., Schmitz, M.D.,
and Bowring, S.A., 2005, Reassessing the uranium
decay constants for geochronology using ID-TIMS
U-Pb data: Geochimica et Cosmochimica Acta, v. 70,
p. 426–445, doi:10.1016/j.gca.2005.09.007.
Scholger, R., Mauritsch, H.J., and Brandner, R., 2000, Permian-
Triassic boundary magnetostratigraphy from the southern
Alps (Italy): Earth and Planetary Science Letters, v. 176,
p. 495–508, doi:10.1016/S0012-821X(00)00026-1.
Shen, S., and 21 others, 2011, Calibrating the end-Permian
extinction: Science, v. 334, p. 1367–1372, doi:10.1126
/science.1213454.
Shields, G., and Veizer, J., 2002, Precambrian marine carbon-
ate isotope database: Version 1.1: Geochemistry Geo-
physics Geosystems, v. 3, doi:10.1029/2001GC000266.
Soddy, F., 1913, Radioactivity: Chemical Society Annual
Report on the Progress of Chemistry, v. 10, p. 262–288,
doi:10.1039/ar9131000262.
Steiger, R.H., and Jäger, E., 1977, Subcommission on geo-
chronology: Convention on the use of decay constants
in geo- and cosmochronology: Earth and Planetary Sci-
ence Letters, v. 1, p. 369–371.
Szurlies, M., 2013, Late Permian (Zechstein) magneto-
stratigraphy in Western and Central Europe, in Gas-
iewicz, A., Roscher, M., and Slowakiewicz, M., eds.,
Late Palaeozoic Climate Cycles: Their Evolutionary,
Sedimentological and Economic Impact: Geological
Society of London Special Publication (in press).
Szurlies, M., Bachmann, G.H., Menning, M., Nowaczyk,
N.R., and Kading, K.C., 2003, Magnetostratigraphy
and high-resolution lithostratigraphy of the Permian-
Triassic boundary interval in central Germany: Earth
and Planetary Science Letters, v. 212, p. 263–278,
doi:10.1016/S0012-821X(03)00288-7.
Szurlies, M., Geluk, M.C., Krijgsman, W., and Kurschner,
W.M., 2012, The continental Permian-Triassic bound-
ary in the Netherlands: Implications for the geomagnetic
time scale: Earth and Planetary Science Letters, v. 317–
318, p. 165–176, doi:10.1016/j.epsl.2011.11.043.
Thomson, J.J., 1913, Rays of positive electricity: Proceed-
ings of the Royal Society of London, ser. A, v. 89,
p. 1–20, doi:10.1098/rspa.1913.0057.
Thomson, W., 1865, The doctrine of uniformity in geology
briefl y refuted: Proceedings of the Royal Society of
Edinburgh, v. 5, p. 512–513.
Torreson, O.W., Murphy, T., and Graham, J.W., 1949, Mag-
netic polarization of sedimentary rocks and the Earth’s
magnetic history: Journal of Geophysical Research,
v. 54, p. 111–129, doi:10.1029/JZ054i002p00111.
Tsukui, K., and Clyde, W.C., 2012, Fine-tuning the calibra-
tion of the early to middle Eocene geomagnetic polar-
ity time scale: Paleomagnetism of radioisotopically
dated tuffs from Laramide foreland basins: Geologi-
cal Society of America Bulletin, v. 124, p. 870–885,
doi:10.1130/B30545.1.
Veizer, J., 1989, Strontium isotopes in seawater through
time: Annual Review of Earth and Planetary Sci-
ence Letters, v. 17, p. 141–167, doi:10.1146/annurev
.ea.17.050189.001041.
Veizer, J., Ala, D., Azmy, K., Bruckschen, P., Bruhn, F.,
Buhl, D., Carden, G., Diener, A., Ebneth, S., Goddris,
Y., Jasper , T., Korte, C., Pawellek, F., Podlaha, O.G.,
Strauss, H., 1999, 87Sr/86Sr, δ13C and δ18O evolution
of Phanerozoic seawater: Chemical Geology, v. 161,
p. 59–88, doi:10.1016/S0009-2541(99)00081-9.
Vine, F.J., and Matthews, D.H., 1963, Magnetic anomalies
over oceanic ridges: Nature, v. 199, p. 947–949, doi:
10.1038/199947a0.
Walker, J.D., and Geissman, J.W, 2009, 2009 GSA Geologic
Time Scale: GSA Today, v. 19, no. 4–5, p. 60–61.
Walker, M., and 17 others, 2009, Formal defi nition and dat-
ing of the GSSP (global stratotype section and point)
for the base of the Holocene using the Greenland
NGRIP ice core, and selected auxiliary records: Jour-
nal of Quaternary Science, v. 24, p. 3–17, doi:10.1002
/jqs.1227.
Whiteside, J.H., Olsen, P.E., Kent, D.V., Fowell, S.J., and
Et-Touhami , M., 2007, Synchrony between the CAMP
and the Triassic–Jurassic mass-extinction event?: Palaeo-
geography, Palaeoclimatology, Palaeoecology, v. 244,
p. 345–367, doi:10.1016/j.palaeo.2006.06.035.
Wilson, P.A., Norris, R.D., and Cooper, M.J., 2002, Testing the
Cretaceous greenhouse hypothesis using glassy forami-
niferal calcite from the core of the Turonian tropics on
Demerara Rise: Geology, v. 30, p. 607–610, doi:10.1130
/0091-7613(2002)030<0607:TTCGHU>2.0.CO;2.
Zachos, J., Pagani, M., Sloan, L., Thomas, E., and Billups,
K., 2001, Trends, rhythms, and aberrations in global
climate 65 Ma to Present: Science, v. 292, p. 686, doi:
10.1126/science.1059412.
Zhu, M.Y., Babcock, L.E., and Peng, S.C., 2006, Advances
in Cambrian stratigraphy and paleontology: Integrat-
ing correlation techniques, paleobiology, taphonomy
and paleoenvironmental reconstruction: Palaeoworld,
v. 15, p. 217–222, doi: 10.1016/j.palwor.2006.10.016.
SCIENCE EDITOR: J. BRENDAN MURPHY
MANUSCRIPT RECEIVED 19 APRIL 2012
REVISED MANUSCRIPT RECEIVED 20 SEPTEMBER 2012
MANUSCRIPT ACCEPTED 26 SEPTEMBER 2012
Printed in the USA
on August 30, 2015gsabulletin.gsapubs.orgDownloaded from
Geological Society of America Bulletin
doi: 10.1130/B30712.1 2013;125, no. 3-4;259-272Geological Society of America Bulletin
J.D. Walker, J.W. Geissman, S.A. Bowring and L.E. Babcock
The Geological Society of America Geologic Time Scale
Email alerting services articles cite this article to receive free e-mail alerts when newwww.gsapubs.org/cgi/alertsclick
Subscribe America Bulletin to subscribe to Geological Society ofwww.gsapubs.org/subscriptions/click
Permission request to contact GSAhttp://www.geosociety.org/pubs/copyrt.htm#gsaclick
official positions of the Society.
citizenship, gender, religion, or political viewpoint. Opinions presented in this publication do not reflect
presentation of diverse opinions and positions by scientists worldwide, regardless of their race,
includes a reference to the article's full citation. GSA provides this and other forums for the
the abstracts only of their articles on their own or their organization's Web site providing the posting
to further education and science. This file may not be posted to any Web site, but authors may post
works and to make unlimited copies of items in GSA's journals for noncommercial use in classrooms
requests to GSA, to use a single figure, a single table, and/or a brief paragraph of text in subsequent
their employment. Individual scientists are hereby granted permission, without fees or further
Copyright not claimed on content prepared wholly by U.S. government employees within scope of
Notes
© 2013 Geological Society of America
on August 30, 2015gsabulletin.gsapubs.orgDownloaded from