You are on page 1of 24

U-Th-Pb Geochronology

Daniel J Condon, British Geological Survey, Nottingham, United Kingdom


Martin J Whitehouse, Swedish Museum of Natural History, Stockholm, Sweden
Matthew SA Horstwood, British Geological Survey, Nottingham, United Kingdom
© 2020 Elsevier Inc. All rights reserved.

Introduction 2
The Age Equation, Decay Constants and 238U/235U 3
Age Equations 3
Decay Constants 4
238 235
U/ U 4
Initial Disequilibrium 4
Initial Pb 4
Pb-Loss 5
Materials for U-Th-Pb Dating 5
Determining U/Pb and Pb/Pb Ratios 6
Isotope-Dilution Thermal Ionization Mass Spectrometry (ID-TIMS) 7
Microbeam Methods (LA-ICP-MS and SIMS) 8
Age Interpretations 9
Data Visualization and Concordance 9
Weighted Mean Dates, Single Analysis Dates 10
Mixtures, Radiogenic Vs. Common/Initial (Isochrons) 10
U-Pb Uncertainty 10
Confidence in the Age Interpretations 11
Application of U-Th-Pb 11
Chronology of the Stratigraphic Record and Calibration of the Geological Time Scale 11
Resolution of Metamorphic, Magmatic and Mineralization Processes 13
Crustal Evolution and the Detrital Record 15
U-Pb in Meteorites and Lunar Samples 19
Conclusions 21
References 22
Further Reading 24

Glossary
Archean One of the four geologic eons of Earth’s history, occurring 4000 to 2500 million years ago (4 to 2.5 Ga).
Blank Lead The additional lead that is contributed to the analyses via the sample processing in the laboratory.
Chondrite A stony (i.e. non-metallic) meteorite that has not been modified, by either melting or differentiation of the parent
body considered to have formed through the accretion of various types of dust and small grains in the early Solar System.
Chondrule A round grain found in a chondrite, interpreted as forming as molten or partially molten droplets in space before
being accreted to their parent asteroids.
Common Lead The lead (204Pb, 206Pb, 207Pb, and 208Pb) that occurs in geological materials and the environment, that is not
directly derived from the decay of 238U, 235U, or 232Th.
Concordia Diagram A plot of two parent/daughter ratios against each other where a ‘concordia curve’ reflects the locus of
points where the two decay schemes are of equal age.
Hadean The earliest geologic eon of Earth’s history, occurring 4500 to 4000 million years ago (4.0 to 4.5 Ga).
Isotope ratio mass spectrometry The measurement and determination of the isotopic spectrum for a sample for a given
element, or elements.
Isochron A plot of the radiogenic daughter isotope (e.g., 206Pb) against the parent isotope (e.g., 238U), all normalized to a
stable isotope of the daughter element (e.g., 204Pb), where the slope is proportional to the age.
Isotope Dilution The addition of a known amount of an enriched or synthetic isotope to sample prior to dissolution, so the
amount of sample isotope, of the same element, can be quantified by IRMS.
Laser ablation inductively coupled plasma mass spectrometry A combined system that uses a laser to ablate and sample a
target prior to introduction and analysis in a mass spectrometer.
Large Igneous Province An extremely large and geologically rapid accumulation of igneous rocks, including intrusive and
extrusive types.

Encyclopedia of Geology, 2nd edition https://doi.org/10.1016/B978-0-08-102908-4.00166-1 1


2 U-Th-Pb Geochronology

Permian-Triassic Mass Extinction Approximately 252 million years ago, it was the Earth’s most severe known biological
extinction event, with up to 96% of all marine species and 70% of terrestrial vertebrate species becoming extinct.
Radiogenic Lead The lead (206Pb , 207Pb , and 208Pb ) that occurs in geological materials that is directly derived from the
decay of 238U, 235U, or 232Th.
Secular Equilibrium The situation in which the quantity of a radioactive isotope remains constant because its production
rate, i.e. the number of atoms of the radioactive members of a decay series are proportional to their respective half lives.
Xenocryst An individual foreign mineral crystal or fragment included within an igneous body.

Abbreviations
a Annum (year)
Ga Giga annum (billion years before present)
ICP-MS Inductively coupled plasma mass spectrometry
ka Kiloannum (i.e. 103 years)
IRMS Isotope ratio mass spectrometry
LA-ICP-MS Laser ablation inductively coupled plasma mass spectrometry
LIP Large igneous province
Ma Mega annum (million years before present)
MSWD Mean square weighted deviation
SIMS Secondary ion mass spectrometry
TIMS Thermal ionization mass spectrometry

Introduction

U-Th-Pb geochronology is one of the most widely applied radiometric dating systems in the geochronology ‘toolbox.’ The
widespread geological occurrence of accessory minerals, especially zircon [ZrSiO4], that are ideally suited as U-Th-Pb geochron-
ometers means that U-Th-Pb geochronology can be applied to a wide range of geological regimes to provide age dates that can be
used to constrain rates of magmatic, metamorphic and sedimentary records and processes. Zircon, and other minerals have the
structural attributes that make them highly suited for radio-isotopic dating, in that they include the parent nuclide in the crystal
structure but preferentially exclude the daughter, and are able to remain as a ‘closed system’ through a wide range of post-formation
geological conditions (e.g., weathering, metamorphism). This geological suitability is further strengthened by: (1) the dual decay of
238
U and 235U over geologically meaningful half lives to stable daughters, 206Pb and 207Pb respectively, which provides an internal
check on whether the radiometric system has been ‘open’ to isotopic disturbance since initially set. This is a critical quality control
that other single parent-daughter isotope decay systems lack and historically has proven critical for evaluating the significance of
dates produced; and (2) the system is underpinned by well-determined decay constants known through first principles counting
experiments. Combined, these attributes mean that the U-Pb system is commonly referred to as the gold standard for geochronology
and often serves as a benchmark for inter-calibration for other geochronometers.
Interestingly, the first U-Th-Pb ages pre-date the ‘discovery’ of isotopes in 1913 by Frederik Soddy. In 1906, Ernest Rutherford
proposed that the radioactive transformation of unstable U to the stable element Pb could be used to determine U-Pb dates and, in
1907, Bertrand Boltwood determined U/Pb ratios for 43 U ore samples and published U-Pb dates ranging from 410 to 2200 million
years (Ma). Arthur Holmes built upon Boltwood’s work, developing more sensitive radiochemical methods for the determination of
U and Pb concentrations in geological materials with much lower U and Pb concentrations, including minerals such as zircon and
feldspar. Unlike the other pioneers, Holmes was a geologist and he went on to investigate the geological settings of the dated
samples and established the beginning of a numerically calibrated geologic timescale, the importance of which to modern-day Earth
Science cannot be understated.
In the following decades, pioneering scientists such as Alfred Nier continued to develop the ‘mass spectrometer,’ while making
key advances, such as determining the (variable) isotopic composition of ‘common’ Pb and making accurate 238U/235U determi-
nations. Following the discovery of nuclear fission, the development of technologies for isotope separation would pave the way for
a major development in U-Pb geochronology—the isotope dilution method. In the 1950s a group at the University of Chicago that
included George Tilton, Mark Inghram, and Clair Patterson made major advances, exploiting enriched isotope tracers and mass
spectrometers, that were then a million times more sensitive compared to those developed in the 1920s, to date different mineral
phases in a 1.0 Ga granite (Tilton et al., 1955) and through the Pb composition of young terrestrial materials and meteorites,
estimate the age of the Earth and meteorites at 4.5 Ga (Patterson, 1956). In the 1970s Tom Krogh pioneered the use of teflon vials
for the dissolution of zircon using hydrofluoric acid - greatly reducing the time needed and the blank contribution. Lead blanks were
U-Th-Pb Geochronology 3

reduced further by the development of the ‘two-bottle still’ for sub-boiling distillation of acids by Jim Mattinson. The synthesis and
purification of 205Pb was due to efforts made by Gerald Wasserburg, Tom Krogh, and Randall Parrish to produce the 205Pb that is
used in the current generation of tracers used for ID-TIMS. With increased precision came a need for more rigorous statistical
analyses. This effort was led for decades by Ken Ludwig, whose Isoplot program is used globally (where MS Excel permits) and the
‘open source’ approach lives on. As precision increased, the understanding of zircon behavior improved, and issues such as Pb-loss,
required innovations such as the ‘chemical abrasion’ method (Mattinson, 2005), and the past three decades have seen a continued
effort to refine these methods, reducing the blank levels, improving materials characterization, measurements, and driving
improvement through application.
Innovation in U-Pb geochronology continued in other directions, particularly for in-situ dating. The use of secondary ion mass
spectrometry (SIMS) in geochronology was pioneered in the wake of the Apollo missions that returned lunar samples to Earth in the
early 1970s (Andersen and Hinthorne, 1972), yielding accurate dates around 4 Ga from a variety of Apollo 11, 12 and 14 samples.
Application to terrestrial samples soon followed (e.g. Hinthorne et al., 1979) with promising results. Technical limitations were
overcome with the development of the first large geometry (LG-) SIMS instrument, the Sensitive High Resolution Ion MicroProbe
(SHRIMP), at the Australian National University (Compston et al., 1984). The SHRIMP instrument design was successfully
commercialized and was soon joined in the market by the CAMECA ims1270 (Schuhmacher et al., 1994).
The first application of a laser ablation system coupled to an ICP-MS was demonstrated for trace element analysis by Gray
(1985). Further studies quickly followed, including the capability for U/Pb geochronology (Longerich et al., 1992; Fryer et al.,
1993). Calibration and control of the U/Pb ratio was poor at this time until Hirata and Nesbit (1995) demonstrated it could be
controlled to an extent, by maintaining laser focus during the ablation. Horn et al. (2000) demonstrated the strong relationship
between U/Pb fractionation and ablation duration and thereafter, a number of studies demonstrated consistent control of U/Pb
with application to geochronology and LA-ICP-MS has become an established method in the geochronologists toolkit alongside
SIMS and ID-TIMS.
The modern incarnation of U-Th-Pb geochronology, using the same fundamental principles over the course of the 20th century,
now employs high-precision methods that can determine U/Th/Pb with precisions of 0.05% and high-spatial resolution analytical
methods that can date parts of minerals, and even map the ages within minerals, at the micron scale. With ever increasing precision
and resolution comes the new challenge: How are the uncertainties currently reported verified? Issues such as inter-laboratory bias,
reproducibility and the nature of uncertainty reporting have come to the fore in the past two decades and required the development
of new experiments to address them. In 2003 an international effort (the EARTHTIME Initiative) aimed at ‘improving geochronol-
ogy’ took on many of these issues through a community effort for sharing best practices and standardization in U-Pb geochronol-
ogy, new calibration experiments, tackling the origin of previously-considered known constants (e.g. 238U/235U) and redetermining
them, developing new reference materials, coordinating inter-laboratory agreement exercises, and developing digital tools for data
reduction and storage. All such efforts are ongoing.
While this chapter is about U-Th-Pb geochronology, its primary focus is on U-Pb systematic and applications. This broadly
reflects the usage in the research literature, reflecting some of the inherent limitations of the Th-Pb system.

The Age Equation, Decay Constants and 238U/235U


Age Equations
One of the basic strengths of the U-Th-Pb geochronology system is that there are three independent decay systems, which lead to the
three fundamental parent-daughter age equations:

206
Pb ∗ ¼ 238 U el238 t − 1 , (1)

207
Pb ∗ ¼ 235 U el235 t − 1 , (2)

208
Pb ∗ ¼ 232 Th el232 t − 1 : (3)

The asterisk in these equations denotes radiogenic 206Pb, 207Pb and 208Pb isotopes that is derived from the decay of 238U, 235U and
232
Th respectively, and l238, l235 and l232 are the decay constants (with their respective half-lives, see Table 1). The half-lives are
suitably long-lived that the U-Th-Pb system can date materials across the range of Earth’s evolution, from 0.5 Ma to 4.567 Ga.
A fourth commonly used age equation can be derived by dividing Eq. (2) by Eq. (1):

Table 1 Summary of the half-lives and decay constants for 238U, 235U and 232Th (Steiger and Jäger, 1977).

Parent Stable Daughter Half life Decay constant


238
U 206
Pb (t1/2–4.47  109 a) 9.8485  10−10/year (0.11%)
235
U 207
Pb (t1/2–7.07  108 a) 1.55125  10−10/a (0.14%)
232
Th 208
Pb (t1/2–1.4  1010 a) 4.9475  10−11/a (1%)
4 U-Th-Pb Geochronology

  
207
Pb ∗ =206 Pb ∗ ¼ 235
U=238 U  el235 t − 1 = el238 t − 1 : (4)

This series of equations provides a range of ways to assess the U-Th-Pb age determined or any given sample, by measuring either a
range of elemental isotope parent/daughter ratios or the 207Pb /206Pb . The 207Pb /206Pb system has the advantage of not requiring
determination of the parent/daughter ratio and only a single element isotope ratio measurement is required. Importantly the
multiple decay systems provide the advantage of being able to test for concordance—i.e. the agreement between ages derived from
two or more of the decay systems. Concordance is an internal measure of whether the material being analyzed has been an open
system and/or whether mixed age components have been analyzed (see Section “Mixtures, Radiogenic vs. Common/Initial
(isochrons)”).

Decay Constants
The determination of the parent/daughter and derived ratios is discussed in Section “Materials for U-Th-Pb Dating.” Critical to
determination of accurate ages is knowledge of the decay constant for a particular isotopic decay series. The decay constants for 238U
and 235U have been determined by a number of experiments, however the most commonly used values are from experiments
carried out about 50 years ago by alpha particle counting of high purity 238U and 235U (Jaffey et al., 1971, Table 1). These values
have been scrutinized, through repeated experiments and ‘inter-calibration’ experiments and so are the best-calibrated, long-lived
decay constants for use in geochronology. As such, many of the other radiometric decay systems (e.g. Rb-Sr, Re-Os, Lu-Hf and
K-Ar/40Ar-39Ar) have used inter-calibration experiments to test and refine their accuracy. By using these decay schemes for materials
that have been dated using the U-Pb system, it can be reasonably assumed that results from the different systems should be age-
equivalent.
High-accuracy U-Pb ID-TIMS analysis has provided further information about the l238/l235 ratio. In closed system zircon, where
the 206Pb/238U and 207Pb/235U ages can be assumed to be equivalent and, if the data are sufficiently precise and accurate, the
l238/l235 ratio can be determined with a precision that improves on the l238/l235 determined by counting experiments
(Mattinson, 2010).

238
U/235U
Knowledge of the isotopic composition of U is required for using Eq. (4) above to calculate a 207Pb /206Pb date. For a long time, it
was assumed that 238U/235U was invariant in nature, in part by an inability to resolve any variation, and a value of
238
U/235U ¼ 137.88 was recommended by the IUGS Subcommission for Geochronology (Steiger and Jäger, 1977). Any deviation
from this value would result in inaccurate 207Pb /206Pb dates but also impact all U-Pb dates as the 238U/235U ratio is typically an
assumed value for data reduction algorithms. Recently, high-precision 238U/235U studies have demonstrated some variation in
geological materials, questioning the use of a single value for 238U/235U ratio in all U-Pb geochronology calculations. This was
substantiated by 238U/235U determinations for a range of U-bearing accessory minerals from a range of geological environments
that revealed variation at the 0.05% level with a lower average value of 238U/235U ¼ 137.818  0.045 (Hiess et al., 2012).

Initial Disequilibrium
U-Pb ages can be determined from Eqs. (1)–(4), if the material being dated formed in secular equilibrium with respect to the
intermediate decay products that form during the decay of U to Pb. This is rarely the case and, thus, the impact of initial
disequilibrium needs to be considered for the nuclides in the respective decay series with geologically significant half lives, i.e.
234
U (t1/2 2.45  105 a) and 230Th (t1/2 7.54  104 a) in the 238U decay chain and 230Pa (t1/2 7.54  105 a) in the 235U decay
chain. The degree to which U-Pb dates are impacted by initial disequilibrium depends on the composition of the geological
reservoir from which the material formed (e.g., magma or aqueous fluid) and the partition coefficients for the different elements for
a given mineral. For U-bearing accessory minerals, those that preferentially include Th (e.g., monazite [(Ce,La,Th)PO4]) have an
excess of initial 230Th that will result in an excess of 206Pb such that the 206Pb/238U date is greater than true U-Pb age. Other
minerals, such as zircon, exclude Th and form with a deficit of 230Th and thus will yield 206Pb/238U dates that are too young (by a
maximum of 110 ka). Corrections for such situations can be applied, with propagated uncertainties, given a knowledge of the
partition coefficients and/or the Th/U ratio of both the parental reservoir and mineral at formation. In U-bearing accessory minerals,
234
U/238U is assumed to not be significantly fractionated. However, some carbonates form from fluids that are highly enriched in
234
U, so the correction for disequilibrium can be significant compared to the age (tens of percent) and, therefore, the strategies for
estimating the correction and uncertainty propagation can be very important.

Initial Pb
Quantifying the amount of 206Pb . 207Pb , or 208Pb in a given mineral is further complicated by the incorporation of ‘common’ Pb
in the analyte, either during formation (initial Pb) and/or during processing (blank Pb). Thus, as illustrated in Fig. 2, the isotopes of
Pb (206Pb, 207Pb and 208Pb) measured in a mass spectrometer are a combination of radiogenic and common Pb (blank  initial).
Subtraction of the common Pb can be achieved by various means, ideally through knowledge of the isotopic composition of the
U-Th-Pb Geochronology 5

common Pb component(s) and referencing to 204Pb, which is a wholly common isotope of Pb. Knowledge of the isotopic
composition of Pb in blank can be obtained via direct measurement of ‘total procedural blanks’ and the various components.
Correction for initial Pb typically uses measured 204Pb and knowledge of the 206Pb/204Pb, 207Pb/204Pb and 208Pb/204Pb. The
minerals zircon, monazite, and baddeleyite [ZrO2] incorporate negligible non-radiogenic Pb and, therefore, corrected ages are
largely insensitive to its composition. Other phases (see below) can incorporate moderate to substantial amounts of initial Pb,
requiring a more accurate estimate, either from a co-existing non-radiogenic mineral phase (e.g. K-feldspar [KAlSi3O8]) or modeled
in an iterative way (e.g. using the terrestrial Pb model of Stacey and Kramers, 1975). Alternatively, an inverse ‘concordia diagram’
can be used to regress uncorrected data through assumed common Pb or a 3-D planar approach applied that makes no assumption
about composition (see below).

Pb-Loss
A particular characteristic of zircons, resulting from their robust physical structure, is their propensity to lose radiogenic Pb as a result
of the in-situ decay of U. The production of high-energy alpha particle emission through radioactive decay results in fluid-accessible
pathways within the crystal structure, along which radiogenic-Pb can be lost. Usually caused by a later thermal event, these Pb-loss
episodes produce a decoupling within the 206Pb/238U, 207Pb/235U and 207Pb/206Pb series that increases the measured U/Pb ratio
and so produce ‘apparent’ ages that are younger than the true crystallization age of the zircon, (Fig. 3). Other datable minerals
(monazite, titanite [CaTiSiO5], rutile [TiO2], xenotime [YPO4]) self-anneal, have a stronger geochemical affinity for Pb or otherwise
appear not to lose Pb in this way.

Materials for U-Th-Pb Dating

There are a set of requirements for geological materials to be considered good candidates for U-Th-Pb geochronology. These are:

1. Preferential incorporation of the parent nuclide (U, or Th) and exclusion of the daughter nuclide (Pb). If the daughter is not
excluded at the time of mineral formation, then either a correction for initial Pb or an isochron approach is required;
2. Behavior as a ‘closed-system’ after mineral crystallization during the range of geological conditions subsequently experienced by
the mineral (e.g. low-temperature burial to high-temperature metamorphism), or have the ‘leaking’ of the daughter nuclide be
linked to an understandable physical process (e.g., diffusion).

A number of accessory minerals have crystal structures that include U and Th and exclude Pb during crystallization and, therefore,
are suitable candidates for U-Th-Pb dating:

• Zircon [ZrSiO4]—a common accessory mineral in felsic and (late stage) mafic igneous rocks, felsic eruptive volcanic rocks, and
reworked in sedimentary systems (initial U  1 to >500 ppm; initial Pb <1 ppm, Th/U 0.01–3);
• Monazite [(Ce,La,Nd,Th)(PO4,SiO4)]—a rare phosphate mineral that occurs as an accessory mineral in igneous and metamor-
phic rocks such as granite, pegmatite, schist, and gneiss (initial U 200 to >5000 ppm; initial Pb <5 ppm, Th/U > 10);
• Xenotime [YPO4]—a rare accessory mineral in granitic rocks that also occurs authigenically as an overgrowth on detrital zircon in
sediments (initial U 100–60,000 ppm; initial Pb <5 ppm, Th/U < 1–5);
• Titanite [CaTiSiO5]—an accessory mineral that occurs in granitic and calcium-rich metamorphic rocks (initial U 1–500 ppm;
initial Pb 5–40 ppm, Th/U < 1–20);
• Baddeleyite [ZrO2]—the predominant Zr-bearing mineral in SiO2 under-saturated rocks and the only U-Pb dateable phase in
mafic dikes, for example. Like zircon it incorporates U in similar amounts, but excludes Pb (initial U 50 to >1000 ppm; initial Pb
<1–10 ppm, Th/U < 1).
• Apatite [Ca5(PO4)3(F,Cl,OH)]—a common phosphate mineral in a wide range of lithologies, and can be of igneous, metamor-
phic or hydrothermal origin (initial U 5 to 200 ppm; initial Pb <5–30 ppm, Th/U < 1 to 20) .
• Calcite [CaCO3] and other carbonate minerals—common as a fracture fillings, as a product of diagenetic/metamorphic
reactions, and through direct precipitation as speleothems in caves or as tufa and travertine at the surface (initial
U < 0.01–100 ppm; initial Pb 5 < 1–100 ppm)
• Other U bearing minerals—a number of mineral phases such as Cassiterite [SnO2] and Hematite [Fe2O3] have recently been
exploited using a combination of microbeam and ID-TIMS analyses to define the radiogenic end-members.
• Meteorites, chondrules, and Ca-Al-rich inclusions (CAIs)—some constituent parts of meteorites contain U at the ppm level and
low initial lead and thus are amenable to Pb-Pb dating (see Section “U-Pb in Meteorites and Lunar Samples”).
Prior to any attempt at dating, samples should be characterized and/or imaged by an appropriate technique to highlight internal
textural and geochemical zonation (Fig. 1). Transmitted light microscopy is useful for examining the external morphology of the
crystals, which can tell something about the nature of the zircon (i.e. primary igneous versus detrital grains reworked in a
sedimentary environment), and for the presence of an inherited core and melt or mineral inclusions. Cathodoluminescence
(CL), back-scattered electron (BSE), electron back-scatter diffraction (EBSD), energy dispersive x-ray and electron microprobe
analysis are all helpful methods for such examination (e.g. Hanchar and Miller, 1993; Reddy et al., 2007). Basic optical microscopy
6 U-Th-Pb Geochronology

Fig. 1 Examples of imaging methods applied to zircon. (A) A polyphase zircon from the Amîtsoq gneiss of southern West Greenland in which SIMS analysis (white
ellipses) reveal Eoachean to Neoarchean ages (Whitehouse et al., 1999); large circular pits are subsequent Hf-isotope analyses (Kemp et al., 2019). (B) Transmitted
light and corresponding BSE and CL images for a zircon from the Adirondack Mountains, of New York, USA (Nasdala et al., 2005), clearly illustrating the
complementary nature of the most commonly utilized SEM-based imaging techniques prior to geochronology. (C) Fine oscillatory zoning and comparatively unzoned
overgrowths in zircon revealed by HF etching of polished mount, which preferentially dissolves radiation-damaged parts of the grain (Nemchin and Pidgeon, 1997).
(D) EBSD images of impact generated polycrystalline zircon and shock deformed monazite from the Yarrabubba impact structure, Western Australia (Erickson et al.,
2020). (E) Raman intensity image of an ultra-high temperature metamorphic zircon from the Napier complex, Antarctica, showing (1) moderately radiation damaged
zircon, (2) amorphous zircon and (3) glassy region of zircon composition (Kusiak et al., 2013).

in both transmitted and reflected light can also be useful to highlight the presence of cracks, cores and inclusions, particularly where
they might be intercepted during microsampling by LA-ICP-MS or SIMS. Targeting of analyses using these methods provides
contextual information within which to interpret U-Pb data, thus offering significant benefits that include improved analytical data
(by avoiding poor quality sample material) and more robust interpretations. For detrital provenance studies, elimination of
sampling bias by avoiding direct targeting is a consideration, but at the speed of analysis now commonplace (100s to 1000s
analyses per day) and the computing technology now available (e.g. image recognition and correlation software), a better approach
is to combine both benefits without significantly impeding the fast workflows and volumes of data required for thorough
investigation of a detrital sample.

Determining U/Pb and Pb/Pb Ratios

The main challenge in U-Pb geochronology is determining the parent/daughter (e.g., 206Pb /238U) ratio for a given U-bearing
mineral. Mass spectrometers measure isotope ratios (isotope ratio mass spectrometry, IRMS), both for a single element (e.g.
207
Pb/206Pb) or, in some cases, between elements (e.g., 206Pb/238U). However, the nature of mass spectrometry is such that the
detected isotope ratios do not reflect the composition of the targeted material because the isotope and elemental ratios fractionate at
U-Th-Pb Geochronology 7

various stages of the analysis process due to differences in mass and the chemical behavior (i.e., differences between elements) of the
isotope for a given element. The manner in which the elemental and isotopic fractionation is corrected is dependent on the
analytical method used.

Isotope-Dilution Thermal Ionization Mass Spectrometry (ID-TIMS)


Analysis by ID-TIMS comprises two stages—chemistry (sample dissolution in the presence of ‘tracer’ isotopes, matrix removal and
separation of elements of interest) and then mass spectrometry, which involves the analysis of Pb and U isotopes via ionization
using a thermal source and isotope ratio mass spectrometry. Unless there is total (i.e., 100%) recovery at every stage of chemical
processing, U and Pb will fractionate relative to each other. Similarly, because Pb and U have very different ionization potentials, Pb
and U will fractionate during thermal source ionization, so that their measured intensities will not reflect the U/Pb ratio of the
analyte. This elemental fractionation during chemistry and mass spectrometry is addressed through the use of ‘tracer isotopes’ that
are either synthetic (non-natural) or highly enriched natural isotopes that can be added to the sample during dissolution to then
‘trace’ any elemental fractionation through to the full process from chemical preparation to mass spectrometry. In modern U-Pb
geochronology, these tracers are typically 205Pb and 233U which are both synthetic and do not naturally occur in the sample. If the
amount of tracer isotope added to the sample is known and the isotope ratio mass spectrometry allows determination of the ratio of
tracer isotope/sample isotope (e.g., 205Pb/206Pb), it is possible to determine the concentration of the isotope of interest. By applying
this approach to both the parent and daughter elements, it is possible to determine the concentration of both parent and daughter
nuclide and, therefore, the parent/daughter ratio of the sample analyzed (Fig. 1).
Another concern is the mass dependent fractionation of the isotopes during ionization, where the lighter isotopes are
preferentially ionized and thus the detected signal needs to be corrected to account for this phenomenon. This can be achieved
either by (i) analysis of a ‘reference material’ containing the same element whose isotopic composition is known, or (ii) by a ‘double
spiking’ process where two synthetic or enriched isotopes are added to the sample such that a known ratio is determined as part of
the measurement. When a known ratio is measured, the difference between the measured and ‘true’ value can be used to quantify,
and correct for, the mass dependent fractionation (Fig. 2).
In practice, a mixed 205Pb(202Pb)-233U(235U) tracer is commonly employed, whereby the 233U and 205Pb are not occurring in
the sample such that the other isotopes can be measured relative to the tracer isotopes (Fig. 1). The 202Pb-205Pb and 233U-235U
‘double spikes’ allow for isotopic fractionation to be quantified and corrected for. Knowledge of the 205Pb/233U ratio in the tracer
combined with the U and Pb isotope ratio information allows the parent/daughter ratios (206Pb/238U, 207Pb/235U) to be
determined. As such, in isotope dilution U-Pb geochronology, U/Pbsample ratios are calculated relative to the U/Pbtracer ratio
(Fig. 1), therefore the calibration of the U/Pb tracer controls the accuracy of the sample U/Pb determinations. Calibration of the
U/Pb ratio for the tracer is also performed by isotope dilution IRMS - this time treating the tracer 233U/205Pb as the unknown and
mixing it with a gravimetric reference solution that has a ‘known’ U/Pb ratio. Gravimetric reference solutions are made from taking
pieces of high-purity (>99.999%) U and Pb metal that have a ‘natural’ isotopic composition that are weighed to high-accuracy and
dissolved in a nitric acid solution to make a solution with a U/Pb known by weight. Using the approach outlined above, IRMS of
U and Pb for a mixed tracer/gravimetric reference solution, where the 238U/206Pb ratio is known via weighing, allows the 233U/205Pb

Fig. 2 Schematic figure illustrating the mass spectrum for Pb and U isotopes, showing the radiogenic parents (238U and 235U), the radiogenic daughter isotopes
(206Pb and 207Pb) and initial Pb isotopes (204Pb 206Pb, 207Pb and 208Pb), and the synthetic/enriched tracer isotopes (235U, 233U, 205Pb and 202Pb). Note that the
intensities represented by the peak heights are only relative for isotopes of the same element, the y-axis for the Pb and U peaks are not comparable and are offset by
some factor that is only known by the U/Pbtracer ratio.
8 U-Th-Pb Geochronology

to be determined and traced back to SI units (i.e., the kilogram). Thus through a linked series of experiments: U/Pbsample relative to
U/Pbtracer, U/Pbtracer relative to U/Pbgravimetirc, U/Pbgravimetric determined by weighing; the sample U/Pb can be determined.

Microbeam Methods (LA-ICP-MS and SIMS)


In contrast to the ID-TIMS approach, which requires dissolution of the sample but permits addition of the tracer isotopes, U-Pb
analyses may also be performed using high spatial resolution methods, which allow targeting of analysis to specific areas within
grains, typically at the 5–30 mm scale. Such analyses may be performed either by Secondary Ion Mass Spectrometry (SIMS, also
known as ion microprobe analysis) or Laser Ablation Inductively Coupled Plasma Mass Spectrometry (LA-ICP-MS). In both of these
approaches, the U/Pb ratio of an unknown sample is measured relative to the U/Pb ratio of a standard for which this is known (by
ID-TIMS), a process known as ‘sample-standard bracketing.’
SIMS utilizes a shaped ‘primary’ ion beam to sputter ‘secondary’ ions from the target material under high vacuum conditions.
The sputtering process results in a wide range of atomic and molecular ions and neutral particles with a broad energy spread that are
admitted to the mass spectrometer. In order to resolve the species of interest (including Pb isotopes and various oxides of U), a large-
geometry instrument is required that is capable of operating at a mass resolving power (defined as M/DM) of at least 4500 that, for
example, can adequately resolve 206Pb+ from 178Hf28Si+ for a zircon analysis. The instrumental mass bias for Pb isotopes is minimal
during SIMS analysis, generally much lower than typical analytical uncertainty for the ratio (Stern et al., 2009). In contrast,
measured inter-element ratios U/Pb and Th/Pb reflect substantial differences in ion yields and require calibration against well-
characterized, matrix-matched reference materials, typically co-mounted with the targets and analyzed at regular intervals (Stern and
Amelin, 2003). Calibration schemes for zircon are robust except for high-U targets (>2000 ppm; White and Ireland, 2012).
In compositionally varied minerals such as monazite and xenotime, multiple reference materials are required (Fletcher et al., 2010).
Calibration bias related to crystallographic orientation occurs in baddeleyite (Wingate and Compston, 2000; Li et al., 2010).
Beam penetration during SIMS is generally limited to a few micrometers, hence depth-profiling across growth zones of different
age or inadvertent sampling of hidden inclusions is rare. In addition to spot analysis, the primary ion beam can be rastered to
generate ion maps from which geochronological data can be extracted at a spatial resolution <1–2 mm (Kusiak et al., 2013).
Laser ablation (LA-)ICP-MS uses a pulsed laser beam focused onto the sample surface to remove material in a sequential fashion.
Samples are housed in a sealed chamber at atmospheric pressure but with He gas flowing across the sample surfaces. This sweeps the
ablated material out of the chamber and down a sample transport tube to the ICP-MS for analysis.
Different laser types and mass spectrometers are used for LA-ICP-MS U-Pb geochronology. Lasers can be either solid-state lasers
or excimer gas lasers. Most commonly three ICP-MS types are used for U-Pb geochronology—quadrupole, single-collector sector-
field, and multi-collector sector-field. Each has different benefits with each similarly capable at measuring the U-Pb ratio. Greater
precision can however be achieved with multi-collector than quadrupole systems when quantifying Pb/Pb, improving ability to
assess concordance (see below).
Like SIMS, U/Pb determined by LA-ICP-MS, needs to be calibrated to a well-characterized matrix-matched reference material to
achieve accurate data. This is due to the strong differential ionization behavior of U and Pb in the Ar plasma altering the original
composition of the material. This is rarely controlled to better than 1–2% (2s) when determined over multiple analytical sessions.
207
Pb/206Pb need not be calibrated relative to a matrix-matched reference material or can use a solution calibration approach,
resulting in improved measurement precision compared to U/Pb. Therefore, although Pb isotopes and the 207Pb/206Pb ratio can be
measured most precisely (to <0.5% 2s) using a MC-ICP-MS, the 206Pb/238U ratio is currently limited at the 2% (2s) level for any
LA-ICP-MS platform. Application of certain analytical and methodological measures is likely to improve this in the future to
0.5–1% (2s).
Laser induced elemental fractionation (LIEF) is common to all forms of LA-ICP-MS. Due to loss of critical focus as the laser
ablates into the sample material (Hirata and Nesbit, 1995) and thermal interaction of the laser beam with the sample material
changes, the ablated particle size distribution decreases with depth and the more refractory U remains in the walls and ejector
blanket of the ablation pit (Košler et al., 2005). By assuming that the reference material ablation process is the same for samples,
LIEF can be corrected. However, this is only a first approximation and in detail each material ablates in a different manner that is
dependent upon its structure, chemistry, and optical properties. This variability is hard to predict and calibrate and contributes to
the overall 2% (2s) limit on the uncertainty of LA-ICP-MS U-Pb geochronology.
Many minerals can potentially be dated using LA-ICP-MS. In addition to the widely exploited zircon, some of the other
commonly targeted minerals for U-Pb geochronology by LA-ICP-MS are monazite, xenotime, titanite, apatite, rutile, and carbon-
ates. All have issues with the quality and availability of reference materials and their characterization and all are similarly
constrained at the 2% (2s) uncertainty limit by the differential ionization behavior of U and Pb within the plasma in the presence
of their specific matrix chemistry.
Microbeam and ID-TIMS methods are very complementary techniques and most powerful when applied together. Coupling
investigative, high-spatial resolution determinations to target sub-sampled extractions for high-precision ID-TIMS quantification,
allows greater confidence in isolating discrete phases while quantifying them to high precision. Providing the ‘best of both worlds,’
this combined approach enables confident and rigorous geochronological (and geochemical) investigations.
U-Th-Pb Geochronology 9

Age Interpretations
Data Visualization and Concordance
Irrespective of the method employed to obtain them, U-Pb data are generally represented in a so-called ‘concordia diagram,’ which
combines the two independent decay schemes 235U to 207Pb and 238U to 206Pb. Two essentially similar forms of the concordia
diagram are in common use: (i) the conventional (or Wetherill) concordia, which plots 206Pb /238U against 207Pb /235U (Fig. 3A
and B) and (ii) the inverse concordia in which 207Pb/206Pb is plotted against 238U/206Pb (Fig. 3C). In both forms, the concordia curve
represents the locus of compositions defined by the decay equations for each isotopic system at any given time. Data points that lie
on the concordia curve have identical Pb/Pb and U/Pb ages and, hence, are considered to represent analyses of closed system
materials and thus a geological event of significance. Data points lying off the concordia curve are termed discordant, yielding
different ages from the two decay schemes and are commonly considered to represent systems that have been disturbed (i.e., open-
system) or are mixed growth domains. The precision of the U-Pb data impacts how well concordance can be resolved, and when the
uncertainties in the 238U and 235U decay constants are considered, the concordia curve is in fact a band (Fig. 3B).
In most cases, the isotopic ratios plotted are just those involving radiogenic Pb (i.e. 206Pb /238U, etc.). In the inverse concordia
diagram (Fig. 3C), however, it is common, especially in certain microbeam studies, to plot ratios involving total 207Pb/206Pb, with
ages being derived either by free or constrained regression of a number of data points to yield a concordia intercept corresponding to

Fig. 3 (A). Conventional concordia showing how the different age equations relate to the concordia curve. (B). Detail of conventional concordia plot showing the
concordia curve as a band reflecting the U decay constant uncertainties. (C) Inverse concordia with regression of discordant data to upper intercept ‘common’ Pb
compositions and lower intercept date. (D) Ranked 206Pb/238U plot for a concordant ID-TIMS dataset (Bed 25, Meishan, see Section “Chronology of the Stratigraphic
Record and Calibration of the Geological Time Scale.” Red bars are single 206Pb/238U dates used in the weighted mean calculation (green bar), olive-green bars are
206
Pb/238U dates excluded (considered xenocrystic). Two interpreted ages from the same dataset are presented.
10 U-Th-Pb Geochronology

an age and a y-axis (equivalent to U/Pb ¼ 0) intercept that is the non-radiogenic Pb composition, or projection from a known or
estimated initial Pb composition through a single datum onto concordia (Fig. 3C).
Before the advent of modern methods that minimize discordance, either by better selection for analysis of single grains,
progressive chemical abrasion, or microbeam methods that permit targeting of apparently pristine areas, it was common to obtain
only discordant data from multiple cogenetic populations (e.g. different grain size fractions), a regression through which yielded
‘upper’ and ‘lower’ intercepts with the concordia curve considered to represent discrete geological events. Today, concordant data
from single and sub-grain studies are commonly pooled to yield 206Pb/238U or 207Pb/206Pb weighted mean dates (see below).
A unique feature of the coupled U-Pb decay systems is that Pb isotope data alone may be used to determine ages. The usual
representation of such data is on a plot of 207Pb/204Pb against 206Pb/204Pb in which suites of cogenetic samples are used to define
isochron ages (see below).

Weighted Mean Dates, Single Analysis Dates


The analytical approaches outlined above are mostly focused on dating single minerals, or domains within single crystals. In order
to derive an age for a dataset consisting of a number of minerals (single crystals, fragments, or domains), it is required to consider
the dataset, that has been analyzed, and how it relates to the target of chronometric interest (i.e., eruptive or intrusive igneous rocks,
age of mineral crystallization, etc.). This requires interpretation of a given set of ‘dates’ to derive an age that has geological
significance.
The weighted mean age calculation is one averaging approach which assigns a loading factor (i.e. weight) to the individual dates
according to their precision. If the uncertainties on the individual analyses are approximately equal (as is typical for microbeam U/
Pb data) then the weighted mean uncertainty is proportional to 1/√n. If they are not equal the more precise data points have greater
weight in the averaging. Over the past two decades, the precision of single data points has increased and revealed complexity in data
sets that was previously masked at the lower levels of precision. This manifests itself as ‘over-dispersion’ where the data points do
not satisfy measures of a single population. One commonly used measure is the mean square weighted deviation or MSWD (also
known as the reduced chi-squared statistic), which is a measure of the degree of coherence within a dataset. The MSWD statistic is
most generally used to assess the appropriate propagation of random and systematic uncertainties, and whether excess (geological)
scatter exists. A value 1 indicates that the scatter in the data can be explained by analytical uncertainties alone, values much less
than 1 indicate that analytical uncertainties have been overestimated, and values greater than 1 can indicate either that the
uncertainties have been underestimated or that another source of scatter, often called ‘geological’ scatter is present. The presence
of over-dispersion could suggest a geological signal that requires a more informed interpretation of the dataset, than a simple
weighted mean of the population.
In addition to ‘weighted mean’ dates based upon a single population, it has become common for subsets of data to be
interpreted as representing the geological event of interest (e.g., deposition of an ash bed) and other dates considered to represent
analyses of material related to another event (e.g., pre-eruptive crystallization). For some situations, this is taken to an extreme
where the youngest single date is considered geologically meaningful. This challenges the geochronologist to consider the
robustness of single U/Pb determinations and their associated uncertainty. With respect to high-precision U-Pb ages for magmatic
zircon, new approaches have been developed (Keller et al., 2018) which use zircon saturation calculations to estimate zircon
crystallization distributions as a function of time or temperature, which are in turn used as ‘prior information’, enabling Bayesian
estimates of the magma eruption time (Fig. 3D).

Mixtures, Radiogenic Vs. Common/Initial (Isochrons)


Some minerals incorporate Pb into their crystal structure during crystallization, often in significant quantities. This ‘initial Pb’ masks
the radiogenic ingrowth of Pb from the decay of U. However, over time, radiogenic Pb becomes a progressively more significant
component of the isotopic composition of the Pb within the crystal. This mixture of common and radiogenic Pb can be best
visualized on an inverse U-Pb diagram (Fig. 3C) where an array of points projects up towards a Pb isotope composition reflecting
the initial Pb end-member, and down towards the intercept with the concordia, where the pure radiogenic end-member resides.
A statistically fitted line through this array of points, taking into account the measurement uncertainty of each data point, represents
a ‘regression’ of the data and ideally defines both the age and initial Pb isotope composition (Fig. 3C).

U-Pb Uncertainty
The uncertainty of an age determination should be considered as important as the age itself (Ludwig, 2003), however it is often
given less attention, or even ignored, when dates are considered and final ages quoted. Understanding the uncertainty and the
context within which it is quoted is critical to the appropriate interpretation of data, and this is very much the case when considering
U-Th-Pb data. No date or age should be cited without reference to its uncertainty and the assigned confidence level at which it
is quoted.
When calculating and quoting uncertainties it is important to understand which components represent ‘random’ and ‘system-
atic’ errors. Random errors occur during (replicate) measurements, can be characterized and the level of uncertainty reduced by
increasing the number of replicate measurements. High-n measurements of the same material, therefore, produce small
U-Th-Pb Geochronology 11

measurement uncertainties. Systematic errors remain predictable during analysis and require some form of calibration or a priori
knowledge to correct for them. In ID-TIMS, U-Pb the calibration of the U/Pb tracer (both its U/Pb ratio and isotopic composition) is
a source of systematic error that has to be considered when considering datasets derived using different U/Pb tracers. In microbeam
analysis, the appropriateness of the inter-element calibration can vary with each session, and so the long-term reproducibility of a
validation material is also an important uncertainty to include in the assessment of systematic errors. Decay constant uncertainties
affect all U/Pb dates equally, and thus do not change the relative offset of one U/Pb date from another, but need to be considered
when comparing with other radiometric chronometers. Final age uncertainties should be presented as  X/Y/Z (or  X/Z if
appropriate), where X is the analytical uncertainty, Y is the analytical uncertainty plus the systematic (tracer or elemental
fractionation) calibration uncertainty, and Z is the total uncertainty including X, Y and the decay constant uncertainties. This
permits use of the data with the level of uncertainty that is appropriate to the problem being addressed.

Confidence in the Age Interpretations


Confidence in any data set can be increased by combining it with other data relevant to the interpretation. In this regard, Hf-isotope
or trace element data combined with U-Pb ages provides extra constraint on the possible age interpretations, as does the
petrographic context of the dated mineral grain within the rock and even the zonation pattern within the mineral grain.
Petrochronology (Kylander-Clark et al., 2013) is one area of geochronological science which seeks to harness the power of petrology
in informing the interpretation of geochronological relationships. This can be used in igneous, metamorphic and stratigraphic
applications, linking the date to the mineral and mineral formation process via chemistry and imaging. For many studies, the
geological context provides an additional framework within which the U-Pb geochronological data can be considered as an external
constraint: cross cutting relationships or the law of superposition in stratigraphic settings.

Application of U-Th-Pb

This section highlights a fraction of the applications of U-Th-Pb geochronology in the Earth sciences, focusing on key areas with
some examples.

Chronology of the Stratigraphic Record and Calibration of the Geological Time Scale
As outlined above, the U-Pb system is ideally suited to dating the U-bearing accessory mineral zircon, which is a common accessory
phase in volcanic ash beds that occur throughout the Earth’s stratigraphic record. This prevalence, combined with the robustness
and accuracy of the zircon U-Pb system, has made it the most appropriate tool for the numerical calibration of the stratigraphic
record (Gradstein et al., 2012). High-precision dating of zircon from ash beds underpins the majority of radiometric dates used for
preparation of the most recent version of the Geological Time Scale, with 176 of the 277 radiometric dates used to calibrate the 2012
GTS being based on the U-Pb system. Furthermore, the U-Pb system is also an input for the calibration of the other decay systems
(e.g., 40Ar/39Ar, Re-Os) used for calibration of the stratigraphic record.
In addition to the development and calibration of the geological timescale, the application of high-precision/accuracy U-Pb
geochronology has been critical to testing hypotheses that link the eruption of large igneous provinces (LIP) to mass extinction
events—pinpointing the timing of both the putative cause (emplacement/eruption of a LIP) and the effect (mass extinction) at a
resolution comparable to the rapidity of the events. Ash beds interstratified with sections that record the extinction in the fossil
record are dated via U-Pb (zircon) and the intrusive and extrusive components of the LIP are similarly targeted for zircon U-Pb
geochronology. The ID-TIMS method can yield (weighted mean) 238U-206Pb dates with uncertainties on the order of <0.05%,
33 ka at 66 Ma, and 125 ka at 250 Ma, sufficient resolution to resolve both the record of extinction and the potential causative
processes.
For over two decades, the end Permian extinction and its relationship to the Siberian Traps LIP has served as an exemplar of high-
precision U-Pb geochronology applied to understanding the causative mechanisms of this largest mass extinction event. During this
time, improvements in U-Pb ID-TIMS geochronology included a significant reduction in the laboratory blanks such that single
grains and grain fragments are now dated compared to multi-grain fractions in the 1990s, and single date uncertainties are reduced.
Another major development was the treatment of zircon to effectively deal with the issue of Pb-loss in zircon, moving on from
mechanical abrasion methods to the ‘annealing and leaching’ developed by Mattinson (2005), such that more of the data can be
used with increased confidence. In the seminal study of Bowring et al. (1998), single 238U-206Pb date uncertainties were on the order
of 0.5 to >1 Myr and the derived weighted mean dates had uncertainties of 0.2 Myr. A study published in 2014 by the same
laboratory produced a dataset with single 238U-206Pb date uncertainties on the order of 0.1–0.2 Myr and weighted mean dates with
uncertainties of 0.03–0.09 Myr. As an example of how these developments have impacted understanding of the absolute age of the
boundary, the dates for Bed 25 (which is 18 cm below the base Triassic GSSP) from four studies over nearly two decades of
analytical improvements are shown in Fig. 4. It is interesting to note from Fig. 3D that using the zircon saturation model (Keller
et al., 2018) on the Burgess et al. (2014) dataset yields an eruption age estimate of 251.824  0.072 (2s).
The most recent study constrains the extinction interval at the end of the Permian to a window of 60 ka following the
deposition of Bed 25 in the GSSP section (Burgess et al., 2014). The same ID-TIMS methods have been used to calibrate the
12 U-Th-Pb Geochronology

Fig. 4 Detail of the base Triassic GGSP which occurs with Bed 27 at Meishan, China. Zircon from ash beds above and below the extinction interval have been
dated using U-Pb ID-TIMS methods constraining both the timing and the duration of the mass extinction event. Bed 25 has been targeted in several studies spanning
the development of U-Pb ID-TIMS methods illustrating how those developments have impacted the precision and accuracy of the derived dates. Red bars are single
206
Pb/238U dates used in the weighted mean calculation (gray bar), green bars are 206Pb/238U dates excluded (considered xenocrystic). Each of the Bed 25 studies
single zircon analyses are chemical abrasion except those by Bowring et al. (1998) that used multi- and single-grain analysis, mechanical abrasion for Pb-loss.

emplacement and eruption history of the Siberian Traps LIP (Burgess et al., 2017) that allows for the assessment of the coincidence
of this LIP with the extinction event, demonstrating that ⅔ of the erupted lava pile was emplaced within a  300 ka interval just
prior to the extinction interval, and that the extinction event was coincident (at the tens of thousands of years level) with the onset of
sill emplacement, allowing testing and refinement of hypotheses that link the evolution of Siberian Trap LIP to the mass extinction
(Burgess et al., 2017).
The same analytical improvements that have allowed more precise U-Pb determinations have also allowed smaller amounts of
Pb to be isolated and measured by ID-TIMS methods, meaning that younger materials that have substantially less Pb can now be
dated. The Bishop Tuff, an eruptive unit that formed as the result of a single caldera forming event, occurs near the base of the
Bruhnes-Normal Chron at 0.78 Ma. As such it has been the target of K-Ar and 40Ar/39Ar studies since the late 1950s and over the past
two decades by U-Pb studies, using both microbeam (SIMS) and ID-TIMS methods (Ickert et al., 2015 and references therein). The
aim of this work has been to constrain the age of the magnetic excursion, and understand the magma evolution and zircon
crystallization history. A large dataset (>600 dates) of SIMS U-Pb dates for zircon yield an age range from 712 to 885 ka, and a
smaller dataset (46) of ID-TIMS dates range from 760 to 780 ka, consistent with an eruption age of 773 ka for the Bishop Tuff that is
based upon astronomical tuning and 40Ar/39Ar dating. Importantly, the zircons analyzed for ID-TIMS U-Pb dating typically
contained <1 pg Pb and this has only been made possible by continued efforts to reduce the laboratory blank contribution to
the hundreds of femtograms level that is now routine in a number of ID-TIMS laboratories.
While dating of ash beds via zircon U-Pb is useful when such rocks occur, it is not the only way to use U-Th-Pb geochronology to
constrain stratigraphic ages. Over the past two decades a number of research groups have made a concerted effort to apply U-Th-Pb
methods to date carbonates formed in caves (speleothems), where flowstones and fossil-bearing sediments are both present and
sometimes inter-stratified, such that dating the flowstones can constrain key fossil assemblages. In South Africa a series of carbonate
hosted caves in the region known as the ‘Cradle of Humankind’ are key examples of this setting, and are the richest source of
hominin fossils outside of East Africa, with at least four species of hominin now recognized (Fig. 5) (Woodhead and Pickering,
2012, and references therein). U-Pb dating of flowstones from these cave systems, with well contextualized but disparate cave
U-Th-Pb Geochronology 13

Fig. 5 A summary of the U/Pb carbonate ages from the South African hominin bearing cave sites, showing the samples/ages in their stratigraphic context for each
site and the time ranges for the fossil hominins from each. From Woodhead J and Pickering R (2012) Beyond 500 ka: Progress and prospects in the U Pb chronology
of speleothems, and their application to studies in palaeoclimate, human evolution, biodiversity and tectonics. Chemical Geology 290–299. doi: 10.1016/j.
chemgeo.2012.06.017, see reference for data sources.

stratigraphies, enable the fossil record of hominins to be reconstructed, providing an absolute timeline for the evolution of
humankind (Fig. 5). In addition to U-Pb analyses of the flowstones, U-Th analyses can be used to ascertain that the materials
have remained closed systems for the past 0.5 Ma.

Resolution of Metamorphic, Magmatic and Mineralization Processes


The occurrence of a range of U-bearing accessory minerals in igneous rocks has resulted in U-Pb dates being used to constrain a
range of magmatic processes. As outlined above, the formation of LIPs can be constrained by U-Pb dating of U-bearing accessory
phases that occur in refractory melts where the U and Zr concentrations are high enough to form zircon  baddeleyite. More recent
studies have taken this approach to successfully search for zircon in mid ocean ridge (MOR) gabbros and used the dating of these
gabbros to constrain the rates of spreading and accretion at slow and fast spreading MORs (Lissenberg et al. 2009, Fig. 6).
While zircon U-Pb has been widely used to constrain the age of silicic magmatic rocks (e.g., granitoids) on all continents to
constrain regional histories and paleogeographic reconstructions, the increased precision and resolution of methods now allows
magmatic processes to be characterized at the 104–105 year resolution. Zircon crystallization and growth within a magma occur as
the magma cools and zircon saturates, the temperature of which depends on the major element chemistry and Zr content. As such,
zircon crystallization over prolonged intervals is likely and, if the magma is saturated, it is possible to entrain and preserve older
(xenocrystic) zircon. This range of processes that can influence the zircon age distribution in a given plutonic complex makes
interpreting zircon data in terms of process difficult without additional information (i.e. mineral chemistry, petrographic context).
However, studies of zircon U-Pb ages from a range of plutonic suites, in general, have been used to constrain processes during
magma transfer and emplacement as plutons, revealing that the timescales for emplacement of plutons is on the order of
3–10  106 years (Fig. 7). The same methods applied to zircon in large volume eruptive products of caldera forming ‘super
eruptions’ tend to reveal a simpler story, with limited pre-eruptive zircon residence (105 years, Fig. 7), and petrographic evidence
highlighting the potential importance of injection of mafic material into low-melt fraction crystal mushes with injection reinvigor-
ating the system leading to eruption. This is highlighted in a study by Wotzlaw et al. (2013) of zircon from the Fish Canyon Tuff
(1000 km3 eruptive volume) where ID-TIMS dating of zircon and trace element analyses on the same volumes (zircon was
14 U-Th-Pb Geochronology

14 VE04-13 VE02-05

13.9

13.8
Pb /238U date (Ma)

13.7
z5 z3
e 13.41 Ma 13.68 Ma
13.6 at
half r 200µm 200µm
m/yr
13.5 1.6 m

15.
206

VE02-03
13.4 VE02-05
CH78-10-005
13.3 CH78-10-007
VE4-13
13.2 z1 z6
199.0 201.0 203.0 205.0 207.0 13.25 Ma 13.59 Ma
200µm 200µm
Distance from ridge axis (km)
Fig. 6 Plot of 206Pb/238U date of zircon in gabbros from the Vema Fracture Zone on the Mid-Atlantic Ridge, against distance from the spreading center with the
youngest dates for each sample defining the spreading rate. Cathodoluminescence images of dated zircon showing simple sector zonation, or no zonation. Modified
from Lissenberg CJ et al. (2009) Zircon dating of oceanic crustal accretion. Science, 323(5917): 1048–1050.

Fig. 7 (A) Comparison of ranked single grain U-Pb zircon analyses from the Mount Givens Granodiorite (blue; 12 samples combined from Frazer et al., 2014) and
the Fish Canyon Tuff (red; 1 sample from Wotzlaw et al., 2013). These two rock units are comparable compositionally and volumetrically. (B) Plot of zircon U/Pb date
versus Yb/Dy for zircon from the Fish Canyon Tuff, which are interpreted to reflect changes in the degree of crystallization prior to eruption (Wotzlaw et al., 2013).
(A) Modified from Coleman DS, Mills RD and Zimmerer MJ (2016) The pace of plutonism. Elements 97–102. doi: 10.2113/gselements.12.2.97.

dissolved and the trace elements were separated from the U and Pb) provide petrologic information about the degree of
crystallization, in this case using Yb/Dy as a proxy for the degree of crystallization, that can be used to infer the temporal evolution
of the magma prior to eruption (Fig. 7; Coleman et al., 2016).
Magmatic systems are also important as hosts to a range of economically important ore deposits, eithers a result of primary
igneous process or the associated hydrothermal systems that are established around crustal magma chambers. While the dating of
the magmatic systems with U-bearing phases such as zircon is straightforward, constraining the timing of the hydrothermal systems
is problematic at the resolution required to understand the links between the pulses of magmatism and the mineralization.
Cassiterite (SnO2) is the primary tin ore mineral precipitated from magmatic-hydrothermal fluids exsolved in association with
reduced granitic systems and pegmatite bodies, and importantly it can contain tens of ppm of U, with variable initial Pb. The acid
resistant nature of cassiterite has impeded its use as a U-Pb chronometer at the potential levels of precision afforded by ID-TIMS
U-Th-Pb Geochronology 15

methods, however recent efforts have developed procedures for total decomposition with low Pb blanks (Tapster and Bright, 2020).
In a demonstration of the accuracy of U-Pb in cassiterite, crystals from the Permian Sn-W greisen deposit at Cligga Head in
southwestern England were dated relative to CA-ID-TIMS zircon U-Pb ages for granites that both precede and post-date Sn-W
mineralization (Tapster and Bright, 2020, Fig. 8). Inverse isochrons U-Pb ages yielding variable compositions of initial common Pb
between crystals demonstrate that robust dates can be extracted and infer the causative magmatic event of ore formation.
In orogenic systems U-Pb geochronology can be used to constrain a range of magmatic and metamorphic processes. The rates of
exhumation of ultra high-pressure (UHP) terranes has been a target for geochronology studies, analyzing zircon and other
U-bearing phases that occur as part of the peak (eclogite) metamorphic assemblage. One example, the Pliocene UHT terrane of
Papua New Guinea, has eclogite rocks that were subducted to pressures  27–31 kbar (100 km depth) and  715  C. Individual
zircons containing inclusions of the peak UHP mineral assemblage yielded ID-TIMS 206Pb/238U dates of 6.0  0.2 Ma to
5.2  0.3 Ma (DesOrmeau et al., 2017). These results constrain the youngest UHP metamorphism on Earth and, combined with
other U-Pb dates that constrain the post-UHP decompression path, provide compelling evidence of rapid (>2.3 cm/yr) exhumation
of the UHP terrane, consistent with data from other orogenic systems.
Deformation of the upper crust represents a challenge for dating, however the common occurrence of calcite as a phase
associated with brittle deformation (faults, fractures and veining) affords an opportunity for U-Pb dating. While the U-Pb dating
of carbonate in cave systems has proved successful, often aided by processes that concentrate U to several to tens of ppm, vein calcite
associated with faults and fractures tends to have U concentrations <1 ppm, making it a greater analytical challenge to isolate the
radiogenic and common end-member domains needed to define an inverse isochron and derive a U-Pb age. The application of
LA-ICP-MS has been instrumental in opening this field of application (Roberts et al., 2020), allowing for the efficient analysis of
samples in order to isolate the radiogenic and common domains, which occur at a scale that were difficult to sample using micro-
mechanical methods (using micro drills or cutting tools). This approach has been applied to a number of systems, often producing
U-Pb dates with uncertainties on the order 4–10% (including the calibration uncertainty) however for events that, by their nature,
have large uncertainties such that even dates at this level of precision are extremely useful. Fig. 9 illustrates an example of U-Pb
dating of syntectonic calcite precipitates that constrain the onset and evolution of faulting in the Dead Sea Transform system (Nuriel
et al., 2017).

Crustal Evolution and the Detrital Record


Many of the minerals commonly used for U-Pb geochronology are physically and chemically robust during weathering, transport,
and sedimentation and, as such, can be used in provenance studies, although zircon is by far the most commonly used (Fedo et al.,
2003). Indeed, the process of sedimentary transport tends to winnow out grains that have been radiation damaged, resulting in a
higher degree of concordance, on average, in detrital populations than in host rocks. Statistical approaches to the interpretation of
detrital zircon suites depend largely upon the purpose of a specific study. If the primary goal is merely to identify the input of a
specific age signature into a sedimentary environment, then sufficient analyses are needed until several grains of this age are

Fig. 8 Inverse isochron plot of U-Pb ID-TIMS data for cassiterite associated with Sn-W mineralization using a low-blank total digestion method. The age of the
cassiterite is independently constrained. Reproduced from Tapster SR and Bright JWG (2020) High-precision ID-TIMS Cassiterite U-Pb systematics using a low-
contamination hydrothermal decomposition: Implications for LA-ICP-MS and ore deposit geochronology. GChron (Geochronology Discussions), doi: 10.5194/gchron-
2019-22.
16 U-Th-Pb Geochronology

Fig. 9 U-Pb Tera-Wasserburg concordia plots for samples from the Shelomo fault zone, Israel, part of the Dead Sea Transform system. Locations of laser ablation
spot analyses (red circles) are indicated on cross-polarized microscopy images. From Nuriel P, Weinberger R, Kylander-Clark ARC, Hacker BR, Craddock JP (2017).
The onset of the Dead Sea transform based on calcite age-strain analyses. Geology 45(7): 587–590.

encountered. More commonly, the goal is to characterize the source region of a detrital zircon population and in this case, the
number of analyses needed is greater. For example, in order to be 95% confident that a single population present at the 5% level has
been sampled requires analysis of at least 59 grains (using the equation of Dodson et al., 1988), while to be 95% confident that all
populations present at the 5% level have been sampled requires at least 117 analyses (Vermeesch, 2004).
In order to yield meaningful information, age data from detrital populations must be filtered to eliminate discordant analyses.
Typically a > 5% discordance filter is applied (Nemchin and Cawood, 2005), although ideally this threshold should take into
account the size of the uncertainty and whether an analysis is concordant within this uncertainty. It is also necessary to choose
whether to use the 206Pb/238U age or 207Pb/206Pb age in such data analysis, the former being more precisely determined after
approximately 1 Ga, the latter being more precise before this time (Fig. 10) when considering typical levels of uncertainty from
microbeam methods. After these initial tests, data are then commonly presented on either probability density plots (PDP), which
take into account individual analytical uncertainties, or kernel density (KD) plots that ignore specific uncertainties, instead using a

Fig. 10 Relationship between uncertainty on measured ratio and uncertainty on derived age, both in percent (%). Due to the long half-life of 238U, uncertainty on
206
Pb/238U ages closely mirrors that on measured ratios; in contrast, the shorter half-life on 235U results in marked magnification of 207Pb/206Pb age relative to
analytical uncertainty for younger targets.
U-Th-Pb Geochronology 17

binning interval. Detrital zircon signatures can be compared by simple stacking of PDP or KD plots in stratigraphic order, from
which the appearance or disappearance of a given peak(s) may be inferred to record a change in provenance at a given stratigraphic
level, for example due to disruption of established drainage patterns or a tectonically significant event in the source region, for
example the docking of an exotic terrane (Pease et al., 2015; Fig. 11).
Detrital zircons are extensively used in modern crustal growth studies, taking advantage of the mineral’s ability to record and
retain several key isotope geochemical signatures pertaining at the time of its crystallization such as Hf, O, and Si isotopes.
In particular, the in situ analysis of radiogenic Hf isotopes, which in zircon are very little changed by in situ Lu decay due to its
low Lu/Hf ratio, are used together with the age of the grain, and its 18O/16O ratio (i.e. d18O value) to define episodes of juvenile
crustal growth (e.g. Kemp et al., 2006), or reworking of older crust (e.g. Næraa et al., 2012) on both a local basis and in estimates of
global continental crustal volume through geological time (Belousova et al., 2010; Dhuime et al. 2012; Fig. 12).
Zircon provides the only remaining material from the Hadean Earth. The discovery (Froude et al., 1983; Compston and Pidgeon,
1986) of a substantial population of >4.0 Ga zircon grains in the Jack Hills conglomerate of Western Australia has resulted in a suite
of (largely) in situ geochemical and isotopic studies aimed at placing constraints on aspects of early Earth evolution (inferring the
presence of surface water on Earth, the nature of P-T regimes, prevailing tectonic processes, and possible traces of biogenic carbon).
All of these studies rely on U-Pb geochronology in zircon to place them in a temporal context over the 400 million years of Earth
evolution represented before the first appearance of a preserved rock record. The majority of Jack Hills zircon grains have been dated
by SIMS (Holden et al., 2009; Whitehouse et al., 2017). Accurate dating is especially critical for interpretation of time-dependent
Hf-isotope evolution and the consequences of inappropriate age assignment were well-illustrated by the earliest Hf-isotope studies
(Harrison et al., 2005), which proposed very early separation of an evolved, continental crust-like reservoir to account for high
initial eHf values. This interpretation was shown to be incorrect when Hf isotopes were measured with concurrent determination of
207
Pb/206Pb age (Kemp et al., 2010), allowing for the filtering of the dataset to include only analyses that had a coherent age

Fig. 11 Probability density plots of concordant U-Pb detrital zircon ages from Permian (A) and Devonian (B) sandstone samples from the New Siberian Islands of
Arctic Russia. The Devonian sample has zircon ages that overlap the uniquely characteristic range for the Baltica basement and the Caledonian Orogen (pink vertical
bars), showing that it was deposited proximal to Baltica. The Permian sample is dominated by zircon ages related to the late-Carboniferous Uralian Orogen (blue
vertical bar) that juxtaposed Baltica and the Siberian craton, with the resulting mountain range cutting off the supply of Baltica-derived sediment. Modified after
Pease VL, Kuzmichev AB and Danukalova MK (2015) The New Siberian Islands and evidence for the continuation of the Uralides, Arctic Russia. Journal of the
Geological Society 172(1): 1–4.
18 U-Th-Pb Geochronology

(A) All (23178)


Number of
200 TC database (13428)
Zircons
Campbell, Allen, 2008 (5246)
Others (4478)

150

100

50

U-Pb Age, Ga

0
0 1 2 3 4
(B)

5 TDMC, Ga

1 Growth curve
PR AR
3
0.8

0.6
2 Integral
Growth
0.4

0.2
1
Age, Ga
0
0 1 2 3 4

0
0 1 2 3 4 Age, Ga 5
(C)
100 Stage 2 Stage 1
(Net growth rate: 0.8 km3.yr–1) (Net growth rate: 3.0 km3.yr–1)
Volume of contentinel crust (%)

Subduction No subduction?

75
inflection

Be II: Crustal reworking


lou corrected
50 so
va
et
al.
20
I: Crustal reworking 10
uncorrected
25

0
0 1 2 3 4
Time since present (Ga)
Fig. 12 Crustal growth models based on a global database of detrital zircon U-Pb ages. (A) Global database of zircon ages from igneous rocks showing distinct
peaks related to supercontinent cycles; (B) Hf isotope based depleted mantle model age of 13,844 dated zircon grains filtered for concordance (Belousova et al.,
2010) and (C) a different approach applied to 6972 dated zircon grains for which O isotopes provide an additional “reworking” constraint on their Hf-based crustal
residence ages (Dhuime et al., 2012).
U-Th-Pb Geochronology 19

Fig. 13 Interpretation of detrital zircon age spectra from the Jack Hills conglomerate, Western Australia. (A) Schematic concordia plot showing differing
crystallization and disturbance ages that may be assigned to a nominally concordant 4000 Ma analysis (inset shows general case for different uncertainty levels).
(B) Histogram of 4880 > 90% concordant ages obtained from 100,000 SIMS analyses of Jack Hills zircon grains (Holden et al., 2009). (C) Results of rigorous filtering
of 451 analyses from 140 Jack Hills zircon grains (Whitehouse et al., 2017) in which only strictly concordant (within 2s uncertainty) analyses have been retained and
further filtered for grains yielding concordia ages from two or more (top panel) and three or more (bottom panel) analyses.

signature. A particular problem in the Hadean (and indeed the Eoarchean and Paleoarchean) is that the concordia curve is
essentially 1-dimensional as a result of the very rapid evolution of 207Pb relative to 206Pb , which, coupled with typical
(1–2%) uncertainty in SIMS-determined U/Pb ratios, means that ancient Pb-loss can result in displacement of an original
composition along concordia to create an ambiguity in age interpretation (Fig. 13). In a study of 146 Jack Hills zircon grains
subjected to multiple in situ geochronological analyses, only a relatively small percentage yielded coherent ages from all analyses
(Whitehouse et al. 2017). Only these analyses were thus interpreted to record true crystallization ages from which Hf- and O-isotope
signatures may be trusted implicitly.

U-Pb in Meteorites and Lunar Samples


The U-Pb system has occupied a pivotal role in our understanding of early Solar System and early planetary processes ever since a
Pb/Pb isochron age of 4.55  0.07 Ga, based on bulk analysis of five meteorite fractions, was published by Patterson (1956).
Subsequent study of meteorites has revealed a degree of complexity in their composition, petrography and evolution that is largely
beyond the ability of the U-Pb system to date with adequate precision, with most chronologies now reliant on relative ages from
short-lived systems such as 26Al-26Mg and 53Mn-53Cr. Nonetheless, these short-lived systems require linking to an absolute time
point. Calcium aluminum-rich inclusions (CAI) in primitive chondrites are considered to represent the earliest high-temperature
condensates in the solar nebula. Advances in analytical methods, particularly application of ID-TIMS step leaching approaches, and
consideration of measured (rather than assumed) U isotopic composition, have now refined the age of CAI’s to 4567.30  0.16 Ma
(Amelin et al., 2010; Connelly et al., 2012), providing a robust time point for the start of Solar System evolution, with precise ages of
chondrules recording a further 3 Ma of remelting prior to establishment of the debris disk from which planetesimals formed
(Fig. 14).
20 U-Th-Pb Geochronology

Fig. 14 Early Solar System chronology and evolution revealed by U-Pb dating of meteorites. (A) reflected light image of the Allende carbonaceous chondrite
showing both CAI and chondrules; (B) Pb-Pb isochron from Allende CAI SJ101 using step-leaching ID-TIMS and measured U isotope composition (Amelin et al.,
2010); (C) revised CAI and chondrule absolute chronology (Connelly et al., 2012).

In the half century since the return of Apollo samples and increasing recognition of lunar meteorites, the U-Pb system has been
extensively applied to the chronology of the Moon, particularly with the advent of in situ SIMS methodologies (see Fig. 15). The
U-bearing minerals zircon and baddeleyite have yielded insights into magmatic processes on the Moon, dating final crystallization
of the post-formation lunar magma ocean (Nemchin et al., 2009) and providing a chronology on subsequent magmatic events
generating Mg-rich and alkaline rocks mafic as well as rare felsic differentiates (Meyer et al., 1996). The highly volatile-depleted
nature of the Moon also renders apatite more readily applicable as a radiogenic mineral, while its lower closure temperatures means
it is more readily reset. Thus, the use of zircon and apatite U-Pb dating in breccias reveals fine structure in both magmatic (Merle
U-Th-Pb Geochronology 21

Fig. 15 Examples of lunar chronology using U-Pb systematics. (A) High precision ID-TIMS dating of zircon from Apollo 14 breccia (Barboni et al., 2017). (B) KD plot
of concordant zircon ages in Apollo 14 breccias showing differences in provenance (Merle et al., 2017); (C) BSE image of Apollo 14 lunar breccia thin section
showing brecciated matrix with dated apatite grain as well as basaltic and impactite clasts (Snape et al., 2016). (D) In situ SIMS dating of Apollo lunar basalt (Snape
et al., 2019); inset shows full data set occupying a ‘mixing triangle’ in inverse Pb-Pb space defined by initial Pb, radiogenic Pb and terrestrial contamination, with the
derived isochron occupying the leftmost side of the triangle.

et al., 2017) and impact processes (Snape et al., 2016). Volcanic glass beads are also highly radiogenic, providing an additional
chronometric material (Nemchin et al., 2011).

Conclusions

U-Th-Pb geochronology is a powerful tool in the geochronology ‘toolbox.’ A wide range of U-bearing minerals occurring in nearly
all geological environments, from the mantle to Earths’ surface, means that U-Th-Pb geochronology is a widely applicable method.
It provides the timeline for most of Earth’s history; from the formation of the Solar System 4.567 billion years ago, the earliest rocks
from 4.35 billions years, the rise of animals and the Cambrian explosion 540 million years ago, through to the evolution of
humankind in the past two million years. U-Th-Pb geochronology applied to crustal rocks (igneous and metamorphic) provides a
means to assess connections between the evolution of the biosphere, climates and the solid Earth.
The absolute nature of the U-Th-Pb system, with well determined decay constants and gravimetrically calibrated U/Pb
determinations, means that U-Pb dates can be considered ‘accurate.’ This means they can be directly compared with other
independent chronometers, such as those based upon astronomical tuning, and they can serve as an anchor for other radio-
isotopic decay systems (e.g., Re-Os, Lu-Hf, 39Ar/40Ar etc.) via inter-calibration experiments, expanding the impact of the U-Pb
system in geochronology.
The development and ongoing refinement of methodologies means that the geoscientist has a range of analytical methods that
can be applied to an increasing range of U- and Th- bearing minerals that form in a range of environments and track a broad range of
22 U-Th-Pb Geochronology

processes. Both microbeam and ID methods are ever increasing in resolution, both spatially and temporally. The nature of the
materials and the questions being asked often dictate the analytical approaches taken. Increasingly, the microbeam and ID-TIMS
methods are applied in combination, using the strengths of both approaches. The increase in data precision/resolution is unveiling
geological complexity that was heretofore masked underneath larger uncertainties. This presents new challenges, ‘standardizing’
data collection and reporting, and objectively interpreting data sets, requiring additional information to guide the age interpreta-
tion. Progress is being made in linking the chronology to the process via the mineral chemistry (‘petrochronology’), improved
process understanding (e.g., zircon saturation in a magma), geological frameworks (e.g., relative chronologies) and information
from other radio-isotopic systems. This is underpinned by community efforts to standardize data collection, ensuring data from
different laboratories are comparable, providing guidance on uncertainty consideration and reporting, and developing new ways for
data (and meta-data) storage, all of which increases the ‘half-life’ of the U-Th-Pb geochronology data.

References
Amelin Y, et al. (2010) U–Pb chronology of the solar System’s oldest solids with variable 238U/235U. Earth and Planetary Science Letters 343–350. https://doi.org/10.1016/j.
epsl.2010.10.015.
Andersen CA and Hinthorne JR (1972) U, Th, Pb and REE abundances and 207Pb/206Pb ages of individual minerals in returned lunar material by ion microprobe mass analysis. Earth
and Planetary Science Letters 14(2): 195–200.
Barboni M, Boehnke P, Keller B, Kohl IE, Schoene B, Young ED, and McKeegan KD (2017) Early formation of the Moon 4.51 billion years ago. Science Advances 3(1): e1602365.
Belousova EA, Kostitsyn YA, Griffin WL, Begg GC, O’Reilly SY, and Pearson NJ (2010) The growth of the continental crust: Constraints from zircon Hf-isotope data. Lithos 119(3–4):
457–466.
Bowring SA, et al. (1998) U/Pb zircon geochronology and tempo of the end-permian mass extinction. Science 280(5366): 1039–1045.
Burgess SD, Bowring S, and Shen S-Z (2014) High-precision timeline for Earth’s most severe extinction. Proceedings of the National Academy of Sciences of the United States of
America 111(9): 3316–3321.
Burgess SD, Muirhead JD, and Bowring SA (2017) Initial pulse of Siberian Traps sills as the trigger of the end-Permian mass extinction. Nature Communications 8(1): 164.
Coleman DS, Mills RD, and Zimmerer MJ (2016) The pace of plutonism. Elements 97–102. https://doi.org/10.2113/gselements.12.2.97.
Compston W and Pidgeon RT (1986) Jack Hills, evidence of more very old detrital zircons in Western Australia. Nature 321(6072): 766–769.
Compston W, Williams IS, and Meyer C (1984) U-Pb geochronology of zircons from lunar breccia 73217 using a sensitive high mass-resolution ion microprobe. Journal of Geophysical
Research: Solid Earth 89(S02): B525–B534.
Connelly JN, Bizzarro M, Krot AN, Nordlund Å, Wielandt D, and Ivanova MA (2012) The absolute chronology and thermal processing of solids in the solar protoplanetary disk. Science
338(6107): 651–655.
DesOrmeau JW, et al. (2017) Rapid time scale of Earth’s youngest known ultrahigh-pressure metamorphic event, Papua New Guinea. Geology 795–798. https://doi.org/10.1130/
g39296.1.
Dhuime B, Hawkesworth CJ, Cawood PA, and Storey CD (2012) A change in the geodynamics of continental growth 3 billion years ago. Science 335(6074): 1334–1336.
Dodson MH, Compston W, Williams IS, and Wilson JF (1988) A search for ancient detrital zircons in Zimbabwean sediments. Journal of the Geological Society 145(6): 977–983.
Erickson TM, Kirkland CL, Timms NE, Cavosie AJ, and Davison TM (2020) Precise radiometric age establishes Yarrabubba, Western Australia, as Earth’s oldest recognised meteorite
impact structure. Nature Communications 11(1): 1–8.
Fedo CM, Sircombe KN, and Rainbird RH (2003) Detrital zircon analysis of the sedimentary record. Reviews in Mineralogy and Geochemistry 53(1): 277–303.
Fletcher IR, McNaughton NJ, Davis WJ, and Rasmussen B (2010) Matrix effects and calibration limitations in ion probe U–Pb and Th–Pb dating of monazite. Chemical Geology
270(1–4): 31–44.
Frazer RE, Coleman DS, and Mills RD (2014) Zircon U-Pb geochronology of the mount givens granodiorite: Implications for the genesis of large volumes of eruptible magma. Journal of
Geophysical Research: Solid Earth 2907–2924. https://doi.org/10.1002/2013jb010716.
Froude DO, Ireland TR, Kinny PD, Williams IS, Compston W, Williams IR, and Myers JS (1983) Ion microprobe identification of 4,100–4,200 Myr-old terrestrial zircons. Nature
304(5927): 616–618.
Fryer BJ, Jackson SE, and Longerich HP (1993) The application of laser ablation microprobe-inductively coupled plasma-mass spectrometry (LAM-ICP-MS) to in situ (U)-Pb
geochronology. Chemical Geology 109: 1–8.
Gradstein FM, Ogg JG, Schmitz MB, and Ogg GM (eds.) (2012) The Geologic Time Scale 2012. Elsevier.
Gray AL (1985) Solid sample introduction by laser ablation for inductively coupled plasma mass spectrometry. The Analyst 110: 551–556.
Hanchar JM and Miller CF (1993) Zircon zonation patterns as revealed by cathodoluminescence and backscattered electron images: Implications for interpretation of complex crustal
histories. Chemical Geology 110(1–3): 1–13.
Harrison TM, Blichert-Toft J, Müller W, Albarede F, Holden P, and Mojzsis SJ (2005) Heterogeneous hadean hafnium: Evidence of continental crust at 4.4 to 4.5 Ga. Science
310(5756): 1947–1950.
Hiess J, Condon DJ, McLean N, and Noble SR (2012) 238U/235U systematics in terrestrial uranium-bearing minerals. Science 335: 1610–1614.
Hinthorne JR, Andersen CA, Conrad RL, and Lovering JF (1979) Single-grain 207Pb206Pb and U/Pb age determinations with a 10-mm spatial resolution using the ion microprobe
mass analyzer (IMMA). Chemical Geology 25(4): 271–303.
Hirata and Nesbit (1995) U-Pb isotope geochronology of zircon: evaluation of the laser probe-inductively coupled plasma mass spectrometry technique. Geochimica et Cosmochimica
Acta 59(12): 2491–2500.
Holden P, Lanc P, Ireland TR, Harrison TM, Foster JJ, and Bruce Z (2009) Mass-spectrometric mining of Hadean zircons by automated SHRIMP multi-collector and single-collector U/
Pb zircon age dating: The first 100,000 grains. International Journal of Mass Spectrometry 286(2–3): 53–63.
Horn I, Rudnick RL, and McDonough WF (2000) Precise elemental and isotope ratio determination by simultaneous solution nebulization and laser ablation-ICP-MS: Application to U-Pb
geochronology. Chemical Geology 164: 281–301.
Ickert RB, et al. (2015) The U–Th–Pb systematics of zircon from the Bishop Tuff: A case study in challenges to high-precision Pb/U geochronology at the millennial scale. Geochimica et
Cosmochimica Acta 88–110. https://doi.org/10.1016/j.gca.2015.07.018.
Jaffey AH, Flynn KF, Glendenin LE, Bentley WC, and Essling AM (1971) Precision measurement of half-lives and specific activities of 235U and 238U. Physical Review Series C
4: 1889–1905.
Keller CB, Schoene B, and Samperton KM (2018) A stochastic sampling approach to zircon eruption age interpretation. Geochemical Perspectives Letters 31–35. https://doi.org/
10.7185/geochemlet.1826.
Kemp AIS, Hawkesworth CJ, Paterson BA, and Kinny PD (2006) Episodic growth of the Gondwana supercontinent from hafnium and oxygen isotopes in zircon. Nature 439(7076):
580–583.
U-Th-Pb Geochronology 23

Kemp AIS, Wilde SA, Hawkesworth CJ, Coath CD, Nemchin A, Pidgeon RT, et al. (2010) Hadean crustal evolution revisited: New constraints from Pb–Hf isotope systematics of the Jack
Hills zircons. Earth and Planetary Science Letters 296(1–2): 45–56.
Kemp AI, Whitehouse MJ, and Vervoort JD (2019) Deciphering the zircon Hf isotope systematics of Eoarchean gneisses from Greenland: Implications for ancient crust-mantle
differentiation and Pb isotope controversies. Geochimica et Cosmochimica Acta 250: 76–97.
Košler J, et al. (2005) Chemical and phase composition of particles produced by laser ablation of silicate glass and zircon—Implications for elemental fractionation during ICP-MS
analysis. Journal of Analytical Atomic Spectrometry 402–409. https://doi.org/10.1039/b416269b.
Kusiak MA, Whitehouse MJ, Wilde SA, Nemchin AA, and Clark C (2013) Mobilization of radiogenic Pb in zircon revealed by ion imaging: Implications for early Earth geochronology.
Geology 41(3): 291–294.
Kylander-Clark ARC, Hacker BR, and Cottle JM (2013) Laser-ablation split-stream ICP petrochronology. Chemical Geology 99–112. https://doi.org/10.1016/j.chemgeo.2013.02.019.
Li QL, Li XH, Liu Y, Tang GQ, Yang JH, and Zhu WG (2010) Precise U–Pb and Pb–Pb dating of Phanerozoic baddeleyite by SIMS with oxygen flooding technique. Journal of Analytical
Atomic Spectrometry 25(7): 1107–1113.
Lissenberg CJ, et al. (2009) Zircon dating of oceanic crustal accretion. Science 323(5917): 1048–1050.
Longerich HP, Wilton DHC, and Fryer BJ (1992) Isotopic and elemental analysis of uraninite concentrates using inductively coupled plasma-mass spectrometry (ICP-MS). Journal of
Geochemical Exploration 43: 111–125.
Ludwig KR (2003) Mathematical-statistical treatment of data and errors for 230Th/U geochronology. In: Bourdon B, Henderson GM, Lundstrom CC, and Turner SP (eds.) Uranium-
series geochemistry. Mineralogical Society of America, Reviews in Mineralogy and Geochemistry, 52: 631–656.
Mattinson JM (2005) Zircon U–Pb chemical abrasion (“CA-TIMS”) method: Combined annealing and multi-step partial dissolution analysis for improved precision and accuracy of
zircon ages. Chemical Geology 220: 47–56.
Mattinson JM (2010) Analysis of the relative decay constants of 235U and 238U by multi-step CA-TIMS measurements of closed-system natural zircon samples. Chemical Geology
186–198. https://doi.org/10.1016/j.chemgeo.2010.05.007.
Merle RE, Nemchin AA, Whitehouse MJ, Pidgeon RT, Grange ML, Snape JF, and Thiessen F (2017) Origin and transportation history of lunar breccia 14311. Meteoritics & Planetary
Science 52(5): 842–858.
Meyer C, Williams IS, and Compston W (1996) Uranium-lead ages for lunar zircons: Evidence for a prolonged period of granophyre formation from 4.32 to 3.88 Ga. Meteoritics &
Planetary Science 31(3): 370–387.
Mundil R, et al. (2004) Age and timing of the Permian mass extinctions: U/Pb dating of closed-system zircons. Science 305(5691): 1760–1763.
Næraa T, Scherstén A, Rosing MT, Kemp AIS, Hoffmann JE, Kokfelt TF, and Whitehouse MJ (2012) Hafnium isotope evidence for a transition in the dynamics of continental growth 3.2
Gyr ago. Nature 485(7400): 627–630.
Nasdala L, Hanchar JM, Kronz A, and Whitehouse MJ (2005) Long-term stability of alpha particle damage in natural zircon. Chemical Geology 220(1–2): 83–103.
Nemchin AA and Cawood PA (2005) Discordance of the U–Pb system in detrital zircons: Implication for provenance studies of sedimentary rocks. Sedimentary Geology 182(1–4):
143–162.
Nemchin AA and Pidgeon RT (1997) Evolution of the Darling range batholith, Yilgarn craton, western Australia: A SHRIMP zircon study. Journal of Petrology 38(5): 625–649.
Nemchin A, Timms N, Pidgeon R, Geisler T, Reddy S, and Meyer C (2009) Timing of crystallization of the lunar magma ocean constrained by the oldest zircon. Nature Geoscience 2(2):
133–136.
Nemchin AA, Whitehouse MJ, Grange ML, and Muhling JR (2011) On the elusive isotopic composition of lunar Pb. Geochimica et Cosmochimica Acta 75(10): 2940–2964.
Nuriel P, Weinberger R, Kylander-Clark ARC, Hacker BR, and Craddock JP (2017) The onset of the Dead Sea transform based on calcite age-strain analyses. Geology (2017) 45(7):
587–590.
Patterson C (1956) Age of meteorites and the earth. Geochimica et Cosmochimica Acta 10(4): 230–237.
Pease VL, Kuzmichev AB, and Danukalova MK (2015) The New Siberian Islands and evidence for the continuation of the Uralides, Arctic Russia. Journal of the Geological Society
172(1): 1–4.
Reddy SM, Timms NE, Pantleon W, and Trimby P (2007) Quantitative characterization of plastic deformation of zircon and geological implications. Contributions to Mineralogy and
Petrology 153(6): 625–645.
Roberts NMW, et al. (2020) Laser ablation inductively coupled plasma mass spectrometry (LA-ICP-MS) U–Pb carbonate geochronology: Strategies, progress, and limitations.
Geochronology 33–61. https://doi.org/10.5194/gchron-2-33-2020.
Schuhmacher M, de Chambost E, McKeegan KD, Harrison TM, and Migeon H (1994) In-situ dating of zircon with the CAMECA IMS-1270. In: Benninghoven A, et al. (ed.) Secondary Ion
Mass Spectrometry, SIMS IX, pp. 919–922. Wiley.
Shen S-Z, et al. (2011) Calibrating the end-Permian mass extinction. Science 334(6061): 1367–1372.
Snape JF, Nemchin AA, Grange ML, Bellucci JJ, Thiessen F, and Whitehouse MJ (2016) Phosphate ages in Apollo 14 breccias: Resolving multiple impact events with high precision
U–Pb SIMS analyses. Geochimica et Cosmochimica Acta 174: 13–29.
Snape JF, Nemchin AA, Whitehouse MJ, Merle RE, Hopkinson T, and Anand M (2019) The timing of basaltic volcanism at the Apollo landing sites. Geochimica et Cosmochimica Acta
266: 29–53.
Stacey JT and Kramers J (1975) Approximation of terrestrial lead isotope evolution by a two-stage model. Earth and Planetary Science Letters 26(2): 207–221.
Steiger RH and Jäger E (1977) Subcommission on geochronology: Convention on the use of decay constants in geo- and cosmochronology. Earth and Planetary Science Letters
359–362. https://doi.org/10.1016/0012-821x(77)90060-7.
Stern RA and Amelin Y (2003) Assessment of errors in SIMS zircon U–Pb geochronology using a natural zircon standard and NIST SRM 610 glass. Chemical Geology 197(1–4):
111–142.
Stern RA, Bodorkos S, Kamo SL, Hickman AH, and Corfu F (2009) Measurement of SIMS instrumental mass fractionation of Pb isotopes during zircon dating. Geostandards and
Geoanalytical Research 33(2): 145–168.
Tapster SR and Bright JWG (2020) High-precision ID-TIMS Cassiterite U-Pb systematics using a low-contamination hydrothermal decomposition: Implications for LA-ICP-MS and ore
deposit geochronology. GChron (Geochronology Discussions). https://doi.org/10.5194/gchron-2019-22.
Tilton GR, Patterson C, Brown H, Inghram M, Hayden R, Hess D, and Larsen E (1955) Isotopic composition and distribution of lead, uranium, and thorium in a Precambrian granite.
Geological Society of America Bulletin 66: 1131–1148.
Vermeesch P (2004) How many grains are needed for a provenance study? Earth and Planetary Science Letters 224(3–4): 441–451.
White LT and Ireland TR (2012) High-uranium matrix effect in zircon and its implications for SHRIMP U–Pb age determinations. Chemical Geology 306: 78–91.
Whitehouse MJ, Kamber BS, and Moorbath S (1999) Age significance of U–Th–Pb zircon data from early Archaean rocks of West Greenland—A reassessment based on combined
ion-microprobe and imaging studies. Chemical Geology 160(3): 201–224.
Whitehouse MJ, Nemchin AA, and Pidgeon RT (2017) What can Hadean detrital zircon really tell us? A critical evaluation of their geochronology with implications for the interpretation
of oxygen and hafnium isotopes. Gondwana Research 51: 78–91.
Wingate MTD and Compston W (2000) Crystal orientation effects during ion microprobe U–Pb analysis of baddeleyite. Chemical Geology 168(1–2): 75–97.
Woodhead J and Pickering R (2012) Beyond 500 ka: Progress and prospects in the U Pb chronology of speleothems, and their application to studies in palaeoclimate, human evolution,
biodiversity and tectonics. Chemical Geology 290–299. https://doi.org/10.1016/j.chemgeo.2012.06.017.
Wotzlaw J-F, et al. (2013) Tracking the evolution of large-volume silicic magma reservoirs from assembly to supereruption. Geology 867–870. https://doi.org/10.1130/g34366.1.
24 U-Th-Pb Geochronology

Further Reading
Allègre CJ (2009) Isotope Geology. Cambridge: Cambridge University Press.
Bowring SA and Erwin DH (1998) A new look at evolutionary rates in deep time: Uniting paleontology and high-precision geochronology. GSA Today 8(9): 1–8.
Dalrymple GB (2001) The age of the earth in the 20th century: A problem (mostly) solved. Geological Society of London Special Publication 190: 205–221.
Dickin AP (2005) Radiogenic isotope geology. Cambridge: Cambridge University Press.
Faure G and Mensing TM (2005) Isotopes-Principles and Applications. New York: Wiley.
Gradstein FM, Ogg JG, Schmitz M, and Ogg G (eds.) (2012) The Geologic Time Scale 2012. Elsevier.
Hanchar JM and Hoskin PWO (eds.) (2003) Zircon: Reviews in Mineralogy and Geochemistry. In: vol. 53. Mineralogical Society of America.
Holmes A (1913) The Age of the Earth. Harper & Brothers.
Holmes A (1947) A revised estimate of the age of the Earth. Nature 159: 127–128.
Holmes A (1947) The construction of a geological time-scale. Transactions of the Geological Society of Glasgow 21: 117–152.
Mattinson JM (2013) The geochronology revolution. Geological Society of America Special Papers 500: 303–320.
Mattinson JM (2013) Revolution and evolution: 100 years of U-Pb geochronology. Elements 9: 53–57.
Nemchin AA, Horstwood MSA, and Whitehouse MJ (2013) High-spatial resolution geochronology. Elements 9: 31–37.
Reiners P, Carlson RW, Renne PR, Cooper K, Granger DE, McLean NM, and Schoene B (2017) Geochronology and Thermochronology. AGU/Wiley.
Schaltegger U, Schmitt AK, and Horstwood MSA (2015) U–Th–Pb zircon geochronology by ID-TIMS, SIMS, and laser ablation ICP-MS: Recipes, interpretations, and opportunities.
Chemical Geology 402: 89–110.
Schoene B (2014) 4.10-U–Th–Pb Geochronology. In: Treatise on Geochemistry. Elsevier.
Schmitz MD and Kuiper KF (2013) High-precision geochronology. Elements 9: 25–30.

You might also like