U-Pb Dating of Mineral Deposits: From Age Constraints To Ore-Forming Processes
U-Pb Dating of Mineral Deposits: From Age Constraints To Ore-Forming Processes
U-Pb Dating of Mineral Deposits: From Age Constraints To Ore-Forming Processes
forming processes
This book chapter has been accepted for publication in: Huston, D. et al. Eds., Isotopes
in economic geology, metallogenesis and exploration. SGA special volume.
U–Pb dating of mineral deposits: from age constraints to ore-
forming processes
1
Univ. Lyon, UJM-Saint-Etienne, UBP, CNRS, IRD, Laboratoire Magmas et Volcans UMR
6524, 42023 Saint Etienne, France
2
School of Earth Sciences, University of Bristol, Wills Memorial Building, Queens Road,
Bristol BS8 1RJ, UK
3
Department of Earth Sciences, University of Geneva, rue des Maraîchers 13, 1205 Geneva,
Switzerland
1. Introduction
The knowledge of the timing and duration of ore-forming processes are perhaps one of the
most desirable pieces of information that geologists require to draw a complete picture of the
deposit and to put its genesis into a coherent regional or even global geological framework. In
many cases, it represents an essential parameter for establishing detailed genetic models, and
can critically impact on exploration strategies. This necessarily requires a reliable, precise and
accurate geochronometer.
In the past two decades, U–Pb dating has seen a remarkable success across the Earth
Sciences to become the most commonly used absolute isotopic geochronometer. This great
success results from considerable improvements in the analytical techniques and in advances
of our understanding of the U–Pb system in the geological environment. The paramount
advantage of U–Pb dating relies on the coexistence of two chemically identical but isotopically
distinct radioisotopes of U (238U and 235U), both of which have their very own decay chain and
decay rates. Furthermore, their half-lives are particularly suitable for geologically relevant ages.
This allows the determination of two independent dates of which equivalence (concordance)
can usually be taken as a sign of the meaningfulness of the date, while discordant dates can be
either geologically irrelevant or may be extrapolated to a meaningful date if the cause(s) of this
discordance can be identified.
The recent success of U–Pb geochronology is the result of numerous stepwise improvements
over the last decades (see detailed history in Davis et al. 2003; Corfu 2013; Mattinson 2013),
but has experienced a boost due to coordinated community efforts (EARTHTIME for isotope
1
dilution analysis: http://www.earth-time.org; PLASMAGE for laser ablation analysis:
http://www.plasmage.org).
Geochronology was born out of the U–Pb system. Radioactivity was discovered at the dawn
of the nineteenth century by H. Becquerel, M. and P. Curie in their work with various uranium
compounds (U-salts, U-metal, pitchblende) (Becquerel 1896a; Becquerel 1896b; Curie et al.
1898; Curie and Skolodowska Curie 1898; Skolodowska Curie 1898). Soon after, E. Rutherford
first suggested that the Pb/U ratio of geological materials could be used to date them
(Rutherford 1906). The next year, B. Boltwood applied this method to 43 uranium ore samples
and obtained the first absolute total-U and total-Pb ages ranging from 410 Ma to 2200 Ma
(Boltwood 1907). This revolution conclusively supported the suggestion made by Charles
Darwin half a century prior, that the earth was several hundred million years old, and was about
to provide absolute age calibrations for the geological timescale of A. Holmes (1911; 1913).
However, it was not until the turn of 1930 that the existence of two radioactive U isotopes and
their respective Pb daughter isotopes was recognized in U ores (Rutherford 1929; Aston 1929;
von Grosse 1932), paving the way for modern U–Pb geochronology. Ever since, improvements
in mass spectrometry, laboratory procedures and advances in nuclear physics have permitted
the analysis of increasingly smaller quantities of U and Pb with improved precision and
accuracy. This in turn, enabled a switch from the analysis of U ore minerals, to low-U bearing
minerals such as zircon, titanite and apatite in the second half of the last century (Larsen et al.
1952; Tilton et al. 1955; Webber et al. 1956; Tilton et al. 1957). However, dating still involved
206
multigrain mineral fractions which typically show discordance between Pb/238U and
207
Pb/235U dates, and render their interpretation subjected to debate, assumption and
uncertainty. The 1970s to 1980s period arguably marks the turning point of U–Pb
geochronology. At that time, the development of low blank single grain zircon dating
(Mattinson 1972; Krogh 1973; Krogh and Davis 1975; Lancelot et al. 1976; Michard-Vitrac et
al. 1977; Parrish 1987), air-abrasion techniques (Krogh 1982) and in-situ ion probe dating
(Hinthorne et al. 1979; Hinton and Long 1979; Froude et al. 1983) concurred to routinely
produce concordant U–Pb ages and triggered an expansion in the range of application of U–Pb
dating across various minerals, geological terrains and planetary materials. The 1990s saw the
advent of the chemical abrasion technique (Mattinson 1994) and of laser-ablation inductively
coupled plasma mass spectrometry (Fryer et al. 1993; Horn et al. 2000) that are now common
practices in many laboratories around the world. This is the time when U–Pb dating was
embraced by the Earth Sciences community, and became an essential tool of geological
mapping and mineral exploration. Perhaps as a sign of a mature discipline, the last decade has
2
seen U–Pb practitioners around the world collaborating in a community driven effort to push
precision, accuracy and inter-laboratory reproducibility of dates toward unprecedented limits,
the EARTHTIME initiative (http://www.earth-time.org).
This century of development of U–Pb dating has left us with a powerful tool for ore deposit
studies. While zircon is arguably the most commonly used and understood mineral due to its
robustness and minimal amount of Pb it can incorporate in its lattice during crystallization (so-
called “common” Pb), a number of other U-bearing minerals are amenable to U–Pb dating (e.g.,
titanite, apatite, monazite, xenotime, rutile, baddeleyite, perovskite, columbo-tantalite,
cassiterite, allanite, calcite, etc). While most minerals can date their crystallization, a handful
of them (e.g., apatite, rutile, titanite) actually date their arrival below their respective closure
temperature for the U–Pb system. This diversity of minerals allows a variety of ore deposit
types and related geological processes (magmatic, hydrothermal, metamorphic, sedimentary
and supergene) to be dated. As we write, U–Pb dates have been published on almost the full
spectrum of deposit types and an increasing number of minerals are being tested and improved
for U–Pb geochronology. However, the systematics of the U–Pb system is only really well-
known in zircon and possibly monazite, followed by titanite, apatite, rutile, baddeleyite, and
xenotime.
Geochronology can illuminate the apparent geological chaos at some deposits or districts, as
well as support, refute or generate hypotheses for ore-forming processes. Nevertheless, only in
rare cases does the dated mineral directly date the ore itself (e.g., columbo-tantalite, cassiterite,
uraninite). As examples, zircon from a porphyry stock dates magma intrusion and not the cross-
cutting copper mineralization, titanite in a skarn dates the high temperature metasomatism and
not the deposition of the polymetallic ore at lower temperature. Some minerals may date
magmatic crystallization (e.g., zircon, baddeleyite), or metamorphic reactions (e.g., monazite,
titanite) and some may date their precipitation from hydrothermal fluids (e.g., monazite,
xenotime, calcite, uraninite). In fact, the meaning of any date remains deeply anchored into
proper field observations and sample characterization. Some minerals and dating methods (e.g.,
fission tracks in apatite and zircon, 40Ar/39Ar in micas and K-feldspar, etc) can also record low-
temperature events that that post-date ore formation, allowing a fuller understanding of the
coupled temperature-time evolution of mineral systems.
While U–Pb geochronology has been extensively used to determine the age of geological
events, it remains to current and future generations of scientists to give increasingly more added
value to increasingly more precise and accurate dates, feeding quantitative and numerical
models or ore-forming processes. For example, when combined with numerical models, the
3
duration of magmatic-hydrothermal events or the probability density distribution of a
population of dates may be interpreted in terms magmatic-hydrothermal flux and volume
(Caricchi et al. 2014, Chelle-Michou et al. 2017). This will be a critical step if we want to
uncover the processes at play during ore formation, and provide mineral exploration
professionals with innovative and efficient tools that may help locating a distant or deeply
buried deposit, or that could provide early information on the potential size of the explored
deposit (e.g., Chelle-Michou et al. 2017).
This chapter reviews the basics of the U–Pb geochronology and the most commonly used
dating techniques and minerals while pointing out their respective advantages, weaknesses and
potential pitfalls. Through a series of case studies, we illustrate the various usages of U–Pb
dating for the study of mineral deposits. Admittedly, U–Pb geochronology is a field that is
strongly biased toward the use of zircon and this chapter is not an exception. Nevertheless, we
will also shed light on U–Pb dating applied to less commonly encountered and dated minerals.
4
where the superscript * indicate the amount of radiogenic Pb that has formed since the system
206
closed. If the system has remained closed since the mineral crystallized, the Pb/238U and
207
Pb/235U dates should be identical. Dividing equations (1) and (2) yield a third age equation:
$%6:; $%6:;
9$%' <=9$%' < $%6 ∗ $-7
:; :; "# , /> ?$-7 @ =A5
$%&:; $%&:;
%
= !$%&"# ( = !$-., ( /> ?$-. @=A5 (5)
9$%' <=9$%' <
:; :;
%
This equation has the advantage that the determination of the age does not require measurement
238
of the U isotopes because the present-day U/235U ratio is mostly constant in U-bearing
accessory minerals and equal to 137.818 ± 0.045 (2σ; Hiess et al. 2012). However, in practice,
207
Pb/206Pb dates are relevant only for ages older than ca. 1 Ga (see below). The constancy of
235 235
this ratio and the low abundance of U further allow the measurement of the U to be
neglected, which is common practice in many laboratories.
Figure 1. (A) Decay chains of 238U and 235U with the approximate half-live indicated for each
radionuclide. (B) Cartoon illustrating the difference between a decay chain in secular
equilibrium and one in disequilibrium. tinitial and tA refer to the time immediately after and some
time after mineral crystallization, respectively.
Decay constants for 238U and 235U are by far the most precisely determined ones among those
used in geochronology. Recommended values are those determined by Jaffey et al. (1971) and
are λ238 = 1.55125 ± 0.00166·10-10 a-1 and λ235 = 9.8485 ± 0.0135·10-10 a-1 (2σ) (Schoene 2014).
However, these constants have been suggested to be slightly inaccurate (Schoene et al. 2006;
5
Hiess et al. 2012), but always within their reported 2σ uncertainties. More accurate values may
be available in the future providing further counting experiments are done.
6
has served isotope geochronologists for nearly two decades. However, Isoplot is no longer
being updated for later versions of Microsoft Excel (last versions working on Excel 2010 on
PC and Excel 2004 on Mac). This led the U–Pb community to start developing multiplatform
replacement geochronological applications such as Topsoil and IsoplotR. Topsoil is being
developed as a Java standalone application by the CIRDLE development team (Cyber
Infrastructure Research & Development Lab for the Earth Sciences, College of Charleston,
South Carolina). IsoplotR is a package developed for the R statistical computing and graphics
software environment by P. Vermeesch (University College London, UK) and can currently be
used through the command line in R or as an online RStudio Shiny applet at
http://isoplotr.london-geochron.com. Both programs currently offer limited functionalities, but
future versions are expected to have similar and probably enhanced capabilities compared to
Isoplot. Finally, density plots (probability density function and Kernel density estimates) can
be drafted with Densityplotter (Vermeesch 2012), a standalone Java-based application. It is
noteworthy that all these programs are freely available online.
Figure 2. Classical plots used to present U–Pb geochronological data. (A) Wetherill concordia
plot with one concordant and one discordant analysis shown as example, (B) Tera-Wasserburg
concordia plot with the same analyses, (C) ranked isotopic date plot for synthetic concordant
data together with the corresponding probability density curve. Note that the while the y-axis
is valid for both the data bars and the density curve, the x-axis labelled “relative probability” is
only relevant for the probability density curve. Single spot/grain dates are ranked only to
facilitate the reading of the figure.
7
it has been the most powerful driving force to advance U–Pb dating during the second half of
the 20th century (Corfu 2013). It is now established that discordance can have a number of
origins including: mixing of various age domains, Pb-loss during physical and chemical
changes in the crystal lattice (partially opened system), initial intermediate daughter isotopic
disequilibrium, incorrect or no correction for non-radiogenic Pb, or a combination of these (Fig.
3). Nevertheless, one should keep in mind that the recognition of some dates as being discordant
is intimately tied to the uncertainty of the data. Indeed, low-precision data might appear
perfectly concordant, while high-precision ones would actually reveal otherwise (e.g., Moser
et al. 2009). This means that any method is blind to discordance at a degree that is inferior to
the best age resolution of that method. Below we present the classical causes of discordance
and the most appropriate ways to avoid, mitigate or value them.
Figure 3. Main causes of discordance plotted on (A) Wetherill concordia diagram and (B) on
a Tera-Wasserburg concordia diagram. Discordance of the red ellipses group is caused by either
mixing of two age domains (one at 2704 ± 9 Ma and one at 743 ± 4 Ma) or by Pb-loss of 2704
± 9 Ma minerals at 743 ± 4 Ma. Discordance of the yellow ellipses group is caused by the
presence of common lead in minerals crystallized at 142 ± 13 Ma (Pbc uncorrected data). Insets
shows the possible vectors of discordance.
8
of any dating result. Imagery using transmitted and reflected light together with
cathodoluminescence (CL) and back-scattered electron (BSE) microscopy greatly aids in this
process but is not always definitive. These images can reveal that a mineral grain can be made
up of a sequence of growth zones starting in the center, and mantled by sequential zones towards
the rim, all of which can have distinct U–Pb ages. Bulk (whole grain) dating of such multi-
domain mineral grains could result in discordant dates, if the age differences are sufficiently
large. A similar effect can arise from dating multigrain mineral fractions if they include grains
with different isotopic ages. In the case of a simple two component mixture of two different
age domains, several analyses could plot along a linear array (a so-called discordia line) in
concordia diagrams, of which the lower and upper intercept dates would correspond to the
respective ages of the two components (red ellipses on Fig. 3). However, multicomponent
mixtures may show more scattered distribution or even plot along artificial, and often poorly
correlated discordia arrays of which the upper and lower intercept dates have no geological
significance, therefore inhibiting meaningful interpretation of the data.
In order to avoid problems arising from mixing several age domains, imagery of the minerals
has become a necessary prerequisite to any dating (either in-situ or whole grain) in order to
accurately place the spot of the analysis (for in-situ dating) or to select only those grains (or
grain fragment) that have one age domain (for whole grain dating). However, small cores or
domains with distinct ages can still go unrecognized if they are present below the imaged
surface or have a similar chemistry to the surrounding zones. This effect may be monitored on
the time-resolved signal for in-situ measurements (changing isotopic ratio) but would hinder
the interpretation of whole grain dates.
9
decreasing order of importance (see Corfu 2013; Schoene 2014 and references therein). At the
sample scale, all these processes will result in discordance of the 206Pb/238U and 207Pb/235U dates
if the age difference is large enough. By calculating by a linear regression through a series of
discordant analyses, upper and lower intercepts ages can be reconstructed, corresponding to the
age of crystallization of the mineral and to the age of the Pb-loss event, respectively (Fig. 3).
Multiple Pb-loss events are notoriously difficult to unravel and may present as excess data
scatter or even spurious discordia lines. Furthermore, highly metamict crystal domains may
also experience U loss or U gain that would result in inversely (i.e., above the Wetherill
concordia) or normally discordant data, respectively. In such cases, no age interpretation can
be done. Complete recrystallization of a grain may lead to complete loss of all accumulated
radiogenic Pb and reset the age to zero. The extremely low diffusion constants for Pb and U in
zircon (Cherniak and Watson 2001; Cherniak and Watson 2003; Cherniak et al. 1997) means
that volume diffusion is a very inefficient process to remove radiogenic Pb from an undisturbed
zircon lattice. It is for this reason that cases of U–Pb system survival have been reported in
granulite facies rocks (e.g., Möller et al. 2003; Kelly and Harley 2005; Brandt et al. 2011;
Kröner et al. 2015).
Open-system-related discordance is caused by several distinct processes that cause fast
diffusion pathways in the zircon lattice, and such discordant data may be difficult to interpret.
Features like multiple growth zones, overgrowth rims, dissolution-reprecipitation textures, or
metamorphic recrystallization can be recognized in BSE or CL images (Geisler et al. 2007).
Furthermore, recrystallized domains have distinct trace element compositions that can be
identified by in-situ chemical analysis (Geisler et al. 2007). Pb-loss through fluid leaching of
metamict domains can result in the deposition of minute amounts of ‘exotic’ elements that
normally would not be able to enter the mineral structure (e.g., Fe or Al in zircon; Geisler et al.
2007). Additionally, the degree of metamictization, crystal ordering and ductile crystal
reorientation can be evaluated with Raman spectroscopy, electron backscatter diffraction
(EBSD), and transmission electron microscopy (TEM), respectively. Finally, for the specific
case of zircon, the chemical abrasion technique (Mattinson 2005) has proven to be a powerful
method for removing zircon domains that have suffered Pb-loss due to fission tracks,
metamictization or other fast diffusion pathways.
2.3.3. Common Pb
Common Pb is a generic name for the fraction of Pb that is not radiogenic in origin and
results from a mixture of initial Pb (i.e., Pb incorporated during mineral crystallization) and/or
10
204
Pb contamination (both in nature and in the lab). The measurement of Pb (the only non-
204
radiogenic Pb isotope) undoubtedly pinpoints the presence of common Pb. However, Pb
measurement can be very challenging for low concentrations of common Pb, or may be prone
204
to isobaric interference with Hg, inherent to the LA-ICPMS technique (see analytical
methods). On a Tera-Wasserburg plot, analyses containing common Pb typically display a
linear array of discordant ellipses defining an upper intercept date older than 4.5 Ma which
207
points to the Pb/206Pb common Pb composition on the ordinate axis, and a lower intercept
providing the age of the mineral (2D isochron; Fig. 3b). If 204Pb/206Pb can be measured, it can
be plotted on a third axis and the data regressed to estimate the common Pb composition, the
age of the mineral and to evaluate the relative contributions of common Pb and Pb-loss on the
cause of discordance (3D isochron; Wendt 1984; Ludwig 1998). This approach has been shown
to provide better precision for the common Pb composition than the 2D isochron method
(Amelin and Zaitsev 2002; Schoene and Bowring 2006). Another Pb-correction practice in LA-
ICPMS and SIMS analysis consists of deducing the common Pb correction from measurement
of 208Pb (stable decay product of 232Th) and by assuming concordance of the U and Th systems.
However, these correction methods may result in overcorrection of some data that are
discordant for reasons other than common Pb only. When possible, it is therefore ideal to apply
a more robust correction based on the direct measurement of the sample 204Pb. The Pb isotopic
composition from laboratory contamination (“blank”) is also an important consideration in
high-precision U–Pb geochronology using isotope-dilution TIMS, and is obtained through
repeated measurement of blank aliquots.
The isotopic composition of initial Pb incorporated during the crystallization of a mineral is
best obtained from measurements of cogenetic low-U minerals such as feldspars, galena or
magnetite. Alternatively, initial Pb compositions for a known age may be estimated from bulk
Earth evolution models (Stacey and Kramers 1975). However, this last approach is less reliable
compared to the measurement of a cogenetic low-U mineral (Schmitz and Bowring 2001;
Schoene and Bowring 2006). Finally, for the specific case of zircon where the presence of
common Pb is essentially limited to inclusions, fractures and metamict domains (see §6.1), the
chemical abrasion technique (Mattinson 2005) has proven to be a powerful method for
removing initial Pb from the crystal, leaving only the need for a laboratory blank correction.
11
of U (Fig. 1b). However, elemental fractionation during mineral crystallization or partial
melting would likely disrupt a previously established secular equilibrium (Fig. 1b). This effect
should ideally be accounted for in geochronology. Nevertheless, most intermediate decay
products of the U series have half-lives of many orders of magnitude smaller (microseconds to
years) than the half-lives of U (Ga; Fig. 1a) and potential disequilibrium would have negligible
effect on the U–Pb dates even at the best of current analytical capabilities (i.e., 0.5‰ uncertainty
on the date). However, intermediate daughters 230Th (238U decay chain) and 231Pa (235U decay
chain) have half-lives that are long enough (75.6 ka and 32.8 ka, respectively; Fig. 1a; Robert
et al. 1969; Schärer 1984; Parrish 1990; Cheng et al. 2013) to critically impact on the accuracy
of the calculated date if disequilibrium is not accounted for (Schärer 1984; Parrish 1990;
Anczkiewicz et al. 2001; Amelin and Zaitsev 2002; Schmitt 2007). For example, during
230
monazite crystallization, Th (of which Th) is preferentially incorporated into the crystal
206
lattice compared to U, thus resulting in excess Pb (e.g., Fig. 1b) and in erroneously old
206
Pb/238U dates if the excess 230Th is not accounted for (Figs. 3, 4a). In turn, the Th-uncorrected
207
Pb/206Pb date for the same crystal would be too young (Fig. 4b). Conversely, zircon
230 206
preferentially incorporates U over Th, rendering Th-uncorrected Pb/238U dates typically
207
too young (Fig. 4a). Similarly, the Pb/235U isotopic system is potentially affected by 231Pa
excess as has been reported for zircon (e.g., Anczkiewicz et al. 2001).
The magnitude of the correction that needs to be applied to correct the isotopic dates for
initial 230Th and 231Pa disequilibrium depends on the distribution coefficient of Th/U and Pa/U
between the dated mineral and the liquid from which it crystallized (a melt or an aqueous fluid),
respectively (Schärer 1984). For the 207Pb/206Pb date, it also depends on the age of the mineral
230 231
(Parrish 1990). Figure 4 shows the effect of initial Th and Pa disequilibrium has on the
206
Pb/238U, 207
Pb/206Pb and 207
Pb/235U dates. It shows that for low mineral/liquid distribution
coefficients (DTh/DU<1) date offsets converge to a minimum of -109 ka and -47 ka for the
206
Pb/238U and 207Pb/235U dates, respectively. However, if the distribution coefficients are high
(>1), excess 206Pb/238U and 207Pb/235U dates up to few Ma can be expected. Conversely, Th/U
distribution coefficient <1 causes excess 207Pb/206Pb dates of few ka to ca. 0.5 Ma (depending
on the age of the mineral), and distribution coefficient >1 causes 207Pb/206Pb dates deficit up to
few Ma for Precambrian samples (Fig. 4b).
232
In practice, the Th/U ratio of the mineral is measured as Th/238U or estimated from the
208
measured amount of its stable daughter isotope Pb by assuming concordance of the U–Pb
and Th–Pb dates. For minerals crystallized from a melt, available Th/U mineral-melt
distribution coefficients (Fig. 4a) can then be used to reconstruct the Th/U of the melt needed
12
for the Th-disequilibrium correction (e.g., adopting the values from Tiepolo et al. 2002;
Klemme and Meyer 2003; Prowatke and Klemme 2005; Klemme et al. 2005; Prowatke and
Klemme 2006; Rubatto and Hermann 2007; Stepanov et al. 2012; Beyer et al. 2013;
Chakhmouradian et al. 2013; Stelten et al. 2015). Alternatively, direct measurement of melt
inclusions hosted in the dated mineral, of glass or of whole rock Th/U ratio are also commonly
used. Choosing the most appropriate estimate of the melt Th/U ratio at the time of mineral
crystallization (using partition coefficient or direct measurement on whole rock or melt
inclusions) should be done at the light of all possible information concerning the crystallization
conditions of the dated mineral (e.g., temperature, crystallinity, co-crystallizing Th-bearing
mineral phases, etc; see examples in Wotzlaw et al. 2014; Wotzlaw et al. 2015).
In essence, 230Th- and 231Pa-corrections are based on the assumption that the dated mineral
crystallized from a liquid in secular equilibrium with respect to the U-series. While this might
238 230
be an acceptable assumption for some magmatic systems (at least for U and Th)
(Condomines et al. 2003), it should not be regarded as a rule, especially for hydrothermal
systems in which Th and U have distinct solubilities (Porcelli and Swarzenski 2003; Drake et
al. 2009; Ludwig et al. 2011). Indeed, the contrasted partitioning behavior of U and Th into a
hydrothermal fluid causes isotopic disequilibrium in the fluid (230Th excess or deficit). In cases
where the existence of this fluid is very short (e.g., for magmatic-hydrothermal systems) no
time is given for radiogenic ingrowth in the fluid which would remain out of secular
equilibrium. Finally, the fractionation of U and Th promoted by the crystallization of U- and
Th-bearing hydrothermal minerals may further enhance isotopic disequilibrium. In such cases,
the Th-correction (or Pa) should aim at determining the Th/U ratio of the last medium where
the decay chain was in secular equilibrium before the crystallization of the mineral. This equates
to determining the bulk source (in secular equilibrium) to sink (dated mineral) distribution
coefficient of Th/U, regardless of the intermediate process(es), assuming short transport
timescales and a unique source of U and Th. For example, Chelle-Michou et al. (2015) used the
Th/U ratio of the porphyries (same as for magmatic zircons; Chelle-Michou et al. 2014) to
correct the dates obtained on hydrothermal titanite from the Coroccohuayco skarn deposit. In
this case, the U-series elements (mainly U and Th) were likely sourced from the magma which
was assumed to be in secular equilibrium and transported to the site of deposition by a magmatic
fluid in a short period of time.
13
Figure 4. Excess in (A) 206Pb/238U and (B) 207Pb/206Pb dates due to initial 230Th disequilibrium,
and (C) excess in 207Pb/235U date due to initial 231Pa disequilibrium as a function of Th/U and
Pa/U mineral/liquid distribution coefficients, respectively (after Schärer 1984; Parrish 1990).
Typical ranges of mineral/melt distribution coefficients for commonly dated minerals are
shown for reference.
14
3. A note on Th–Pb geochronology
Although less commonly used than U–Pb geochronology, Th–Pb dating may, in some cases,
be advantageous and complementary to U–Pb dating. Due to comparable ionic radii of U and
Th and similar valence (tetravalent except for oxidized systems where U in mostly hexavalent),
most minerals hosting U into their structure will also incorporate Th (if it is available in the
232
system), and vice versa. The single long-lived isotope of Th, Th, decays to 208Pb through a
chain of alpha and beta decays. The Th–Pb decay offers the possibility of a third independent
geochronometer embedded within the mineral allowing for a further assessment of the
235
robustness and meaningfulness of the obtained date. In addition, the nearby masses of U,
235 232 204 206 207 208
U and Th on one side, and of Pb, Pb, Pb, and Pb on the other side, allows for
simultaneous measurement of U–Th–Pb isotopes from the same volume of analyte (ablated
volume or dissolved grain). The generalized age equation writes as follow:
$%. $%. $-$
"# "# BC
!$%'"# ( = !$%'"# ( + !$%'"# ( /𝑒 1$-$ 2 − 15 (6)
*
232
Where λ232 is the Th decay constant. If common Pb is negligible equation (6) can be
simplified to:
$%.
"#∗
! $-$BC ( = 𝑒 1$-$ 2 − 1 (7)
The 232Th decay constant is much smaller to that of 235U (half-life of 14 Ga) and is commonly
considered to be 4.947 ± 0.042·10-11 a-1 (2σ; Holden 1990). Despite a good accuracy of the
232
Th decay constant as suggested by the common concordance of Th–Pb and U–Pb dates (e.g.,
Paquette and Tiepolo 2007; Li et al. 2010; Huston et al. 2016), its precision is an order of
magnitude lower than those of 238U and 235U. This can represent the main source of systematic
uncertainty on Th–Pb dates and the main limitation of this system when working below the
percent precision level. However, unlike uranium, intermediate daughter isotopes of the 232Th
decay chain have short half-lives such that any isotopic disequilibrium formed during mineral
232
crystallization will fade within few decades only. Therefore, the Th decay chain can be
considered to have remained in secular equilibrium on geological timescale. It results that on
cases where U–Pb dates require a large initial 230Th-disequilibrium correction and parameters
required for this correction are difficult to estimate (e.g., hydrothermal minerals), Th–Pb dates
may be much more accurate than U–Pb ones (but often of lower precision).
232
Due to the very long half-live of Th, the optimal use of Th–Pb geochronology (highest
analytical precision) is achieved for old sample and/or minerals with high Th concentrations so
208
that large amount of Pb have been accumulated. In the case of Th-rich minerals (e.g.,
15
monazite and perovskite, and, to a lesser extent, xenotime, apatite, titanite and allanite),
thorogenic 208Pb (i.e., 208Pb*) would typically be so abundant than common Pb correction may
208
not introduce significant uncertainties into the computed Pb*/232Th ratio or may even be
neglected.
208
Pb/232Th dates are most commonly presented as bars of which the center represents the
date and the length reflect the associated uncertainty. To evaluate the concordance of the Th–
Pb and U–Pb systems, concordia diagrams (208Pb*/232Th vs 206Pb*/238U or 207Pb*/235U) offer a
convenient graphical representation of the data.
16
BSE techniques prior to analysis. Typical analytical uncertainties for zircon dates are on the
order of 3-5 % for single spot and of 0.2-2 % for the weighted mean dates (Fig. 5). However,
accuracy may not be better than 3% (Klötzli et al. 2009; Košler et al. 2013), which should be
considered when comparing LA-ICPMS U–Pb dates from different studies or with dates from
other isotopic systems.
Table 1. Comparison of the three analytical techniques used for U–Pb dating.
The LA-ICPMS setup consists of a laser of short wavelength in the UV range (typically 193
nm), an ablation cell and an ICPMS instrument. The sample is placed into the ablation cell
along with several standards. During ablation, repeated laser pulses are focused on the surface
of the dated mineral. The resulting ablated aerosol is subsequently transported by a carrier gas
(usually He ± Ar ± N2) toward the Ar-sourced plasma torch at the entry of the mass spectrometer
where it is ionized and transferred into the ion optics of the mass spectrometer. LA-ICPMS U–
17
Pb dating is mostly carried out on single-collector sector-field ICP-MS instruments that offer
sequential measurement of individual Pb and U isotopes in a mixed ion-counting – Faraday cup
mode.
The spot size used for LA-ICPMS geochronology mainly depends on target size and the U
concentration of the dated mineral. As a reference, 25-35 µm spots are commonly used for
zircon and can be as low as 5 µm for monazite (Paquette and Tiepolo 2007). Crater depth for a
30-60 s analysis is on the order of 15-40 µm depending on the fluence of the laser and on the
ablated material. However, laser-induced U–Pb fractionation increases with crater depth during
ablation, which negatively impacts on the analytical uncertainty of the measured Pb/U ratio.
Ultimately, this is an important limiting factor for precision and accuracy in LA-ICPMS
geochronology (Košler et al. 2005; Allen and Campbell 2012). The technique requires a laser
setup that yields reproducible ablation with small particles (subsequently more efficiently
ionized in the plasma torch) and that limits crater depth to no more than the spot diameter by
minimizing the laser fluence (e.g., Günther et al. 1997; Horn et al. 2000; Guillong et al. 2003).
Another important limitation of LA-ICPMS U–Pb dating is the imprecise common Pb
204
correction due to the difficulty of precisely measuring common Pb due to an isobaric
204
interference with Hg (traces of Hg are contained in the Ar gas). Common Pb correction
208
protocols using Pb may be employed and are preferred over simple rejection of discordant
analyses. It results that, age interpretation of minerals with elevated common Pb contents (e.g.,
titanite, rutile) may be hampered by large age uncertainties due, in part, to the large uncertainties
associated with the common Pb-correction.
LA-ICPMS and SIMS (see below) U–Pb dating are comparative techniques that require
analysis of a reference material, which is as close as possible to the chemical composition and
the structural state of the unknown (sample). It is analyzed under identical ablation conditions
to the sample to determine the machine fractionation factor of any measured element
concentration; this fractionation factor is then applied to the element ratios and concentrations
of the unknowns. A series of analyses unknown (~10) is typically bracketed by analyses of a
reference material (~2-4) to correct for elemental fractionation and monitor for machine drift.
In addition, at least one secondary standard should be repeatedly analyzed during the same
session in order to demonstrate the accuracy of the fractionation correction. This enables an
estimate of the long-term excess variance of the laboratory that is required in the uncertainty
propagation protocol (see below). A list of commonly used reference materials and their
reference values is provided in Horstwood et al. (2016). Standards for LA-ICPMS and SIMS
U–Pb dating should be homogenous in age, trace element composition, and have comparable
18
trace element concentration and structural state (matrix match) as the unknowns (Košler et al.
2005). Failure to match the matrix of the unknown results in different ablation behavior (rate,
stability, fractionation) and ultimately compromises the accuracy of the date (Klötzli et al.
2009). Therefore, a mineral of unknown age should be standardized using a reference material
from the same mineral. Furthermore, different degrees of metamictization also impact on the
matrix match between standards and unknowns and can be an important source of inaccuracy
for zircon dates (as much as 5 % inaccurate; Allen and Campbell 2012; Marillo-Sialer et al.
2014) and possibly for other minerals as well (e.g., titanite, allanite, columbo-tantalite).
Interlaboratory comparisons for LA-ICPMS and SIMS U–Pb dating have highlighted
discrepancies of U–Pb ages for a series of standards measurements which is sometimes outside
of the reported 2σ uncertainties (Košler et al. 2013). This is thought to reflect different data
reduction strategies in different laboratories (e.g., Fisher et al. 2010) and uncertainty
propagation protocols, that are not always thoroughly documented. This has triggered a
community driven effort to establish standard data reduction workflow, uncertainty propagation
protocols, and data reporting templates (Horstwood et al. 2016) that should be embraced by the
LA-ICPMS community. New community-derived standards for LA-ICPMS dating suggest the
use of the x/y/z/w notation for uncertainty reporting where: x refers to the analytical (or random)
uncertainty, y includes the variability of standards measured in the same lab, z includes the
systematic uncertainty of the primary standard isotopic composition (and of the common Pb
correction if appropriate), and w includes the decay constant uncertainty (Horstwood et al.
2016; McLean et al. 2016). Comparing LA-ICPMS U–Pb data with data from other LA-
ICPMS, SIMS or ID-TIMS laboratories should be done at the z uncertainty level, while
comparison with geochronological data from other isotopic systems have to include decay
constant uncertainties (Chiaradia et al. 2013). Raw data processing, visualization and
uncertainty propagation protocols for LA-ICPMS U–Pb dating have been implemented in the
freely available ET_Redux software (McLean et al. 2016) and allow more robust interlaboratory
data comparison and collaborative science.
19
Typical SIMS craters are 10-15 µm in diameter and 1–2 µm deep, therefore this technique has
higher spatial resolution and is by far less destructive than LA-ICPMS and permit subsequent
isotopic analysis (e.g., O, Hf–Lu) to be done on the same spot (slight repolishing would be
required before SIMS analysis). Analysis is done directly from a thin section, polished grains
mounted in epoxy resin, or from entire grains pressed into indium when analyzing U and Pb
isotopes along a profile from the surface to the interior of a grain (depth profiling). The accuracy
of the obtained result depends on extrinsic factors such as the position of standard and
unknowns in the mount and the quality of the polishing. SIMS analysis of zircon typically yields
U–Pb dates of 0.1–1% precision and accuracy (Fig. 5); it is the preferred method when
analyzing complex minerals (e.g., thin metamorphic rims), very small grains (e.g., xenotime
outgrowths on zircon; McNaughton et al., 1999) or valuable material.
Pb isotopic fractionation in SIMS is subordinate when compared to LA-ICPMS techniques.
Therefore, 207Pb/206Pb dates can be calculated directly from counting statistics. In contrast, there
is a significant difference in the relative sensitivity factors for Pb+ and U+ ions during SIMS
206
analysis. The fractionation of the Pb+/238U+ ratios is highly correlated with simultaneous
254
changes in the UO+/238U+ ratios which forms the basis of a functional relationship that
enables the calibration of the 206Pb/238U dates. Although the 206Pb+/238U+ versus 254
UO+/238U+
238
calibration is the most widely used, other combinations of U+, 254
UO+ and 270
UO2 have
proved successful. As in the case of LA-ICPMS, the SIMS 206Pb/238U calibration is carried out
with reference to a matrix matched reference material (e.g., Black et al. 2004). This is quite
straightforward for zircon and baddeleyite (ZrO2), but more difficult for chemically and
structurally more complex minerals (e.g., phosphates, complex silicates, oxides). In the latter
cases, matrix correction procedures using a suite of reference materials accounting for the effect
of highly variable amount of trace elements have been developed (e.g., Fletcher et al. 2004;
Fletcher et al. 2010). Calibration biases are also introduced through different degrees of
structural damage from radioactive decay (White and Ireland 2012). It is highly recommended
to analyze a reference zircon as unknown again to control the accuracy of the technique
(validation or secondary standard; Schaltegger et al. 2015).
The common Pb correction is carried out via measurement of 204Pb, 207Pb or 208Pb masses.
The main challenge of SIMS analysis is the resolution of molecular interferences on the masses
of interest (Ireland and Williams 2003), which requires careful consideration when analyzing
phosphates or oxides.
No standard data treatment protocol exists for SIMS dates. In fact, the two types of
equipment (SHRIMP from Australian Scientific Instruments and IMS 1280/90 from
20
CAMECA) provide very differently structured data that require different data treatment
software.
Figure 5. Typical analytical uncertainties for zircon 206Pb/238U, 207Pb/235U, 207Pb/206Pb single
spot/grain dates for modern (A) LA-ICPMS, SIMS and, (B) CA-ID-TIMS dating techniques.
Weighted mean dates refers to the weighted mean of a set of statistically equivalent single
spot/grain dates based the most precise isotopic ratio (typically 206Pb/238U for dates younger
than ca. 1 Ga and 207Pb/206Pb for dates older than 1 Ga)
21
4.3. Isotope dilution-thermal ionization mass spectrometry (ID-TIMS)
The U–Pb method that offers the highest precision and accuracy is Chemical Abrasion,
Isotope Dilution, Thermal Ionization Mass Spectrometry (CA-ID-TIMS; Table 1, Fig. 5). This
method involves the dissolution and analysis of entire zircon grains and other accessory
minerals, and, hence, disregards any protracted growth history recorded in this grain. Zircon
imaging prior to dating can be taken to increase the chances of analyzing a single-aged grain or
grain population. The ID-TIMS community is organized as a part of the EARTHTIME
consortium (Bowring et al. 2005), which is working together to improve precision and accuracy
of U–Pb dating.
It is now standard to pre-treat zircons with the “chemical abrasion” procedure of Mattinson
(2005). This process involves heating the zircon at 900°C for 48 hours, followed by partial
dissolution in HF + HNO3 at 180-210°C for 12 to 18 hours. The heating re-establishes the zircon
crystalline structure by annealing any radiation-related structural damage in slightly affected
domains. The partial dissolution procedure then only removes domains with more severe
structural damage and leaves a proportion of the original grain behind. The surviving zircon
fragment is then considered to be perfectly crystalline and is used for isotope ratio analysis.
Chemically abraded zircon grains are recognized to be more concordant and provide more
reproducible U–Pb results. This treatment is not currently applied for SIMS or LA-ICPMS
analysis techniques, but initial experiments have yielded positive results (Kryza et al. 2012;
Crowley et al. 2014; von Quadt et al. 2014). The procedure has been tested on other accessory
phases including baddeleyite (Rioux et al. 2010), but without clear evidence of improving
concordance.
The dissolved grains are mixed with a (202Pb–)205Pb–233U–235U tracer solution (e.g., as
provided by EARTHTIME; ET535 and ET2535; Condon et al. 2015; McLean et al. 2015), and
the Pb and U isotopes isolated from other trace elements through chromatography. Isotopic
compositions are most commonly measured as Pb+ and UO2+ on a thermal ionization mass
spectrometer from the same filament either by ion counting methods (using a secondary
electron multiplier or a Daly-based photomultiplier device), or by a combination of ion counters
and high-sensitivity, high-resistance Faraday collectors. Uranium may also be measured
separately as U+ by solution MC-ICP-MS utilizing a mixed ion counting – Faraday
measurement setup, or as U+ on a double or triple filament assembly in a TIMS.
An important part of high-precision, high-accuracy U–Pb geochronology is the correct
treatment of all sources of uncertainty and their correct propagation into the final age. The ID-
22
TIMS community has been adopting the x/y/z notation for uncertainty reporting (e.g., 35.639 ±
0.011/0.014/0.041 Ma) where: x is the random uncertainty (or analytical; including counting
statistics, common Pb and Th-disequilibrium corrections), y includes the systematic uncertainty
from tracer calibration and, z includes the decay constant uncertainty (Schoene et al. 2006;
Schoene and Bowring 2006; McLean et al. 2011). Comparison of ID-TIMS U–Pb data with U–
Pb data from SIMS or LA-ICPMS techniques should consider the y uncertainty level, while
comparison with data from other isotopic systems (e.g., Re-Os, 40Ar/39Ar) should include both
decay constant and systematic uncertainties (z level). Final age precision is mainly defined by
the ratio of radiogenic to common Pb (Pb*/Pbc), which is, in the case of zircon, a function
mainly of procedural Pb blank. Total blank levels of <0.5 pg of Pb are currently state-of-the-
art.
The EARTHTIME community has generally accepted and adopted a software package
consisting of Tripoli raw data statistics and U-Pb_Redux data treatment and visualization
(Bowring et al. 2011; McLean et al. 2011).
23
Figure 6. Typical range of closure temperature for minerals used for U–Pb dating. Dark grey
bars indicate robust closure estimates while light grey bars indicate approximate estimates.
Modified from Chiaradia et al. (2014), with additional data for apatite (Cochrane et al. 2014),
rutile (Vry and Baker 2006), baddeleyite (Heaman and LeCheminant 2001), garnet (Mezger et
al. 1989), xenotime and allanite (Dahl 1997).
24
Figure 7. Ranked LA-ICPMS and CA-ID-TIMS 206Pb/238U zircon dates and weighted means
for the hornblende-biotite porphyry (sample 10CC51) from the Eocene Coroccohuayco
porphyry-skarn deposit, Peru. Data from Chelle-Michou et al. (2014). Single spot/grain
analyses are plotted at the level of their analytical uncertainties (2σ) and weighted mean dates
include the analytical uncertainties and: (i) an additional excess variance obtained from
repeated measurement of the secondary standard (91500) and the systematic uncertainty in the
standard mineral isotopic composition, for LA-ICPMS data; (ii) the systematic uncertainty
related to the composition of the isotopic tracer, for CA-ID-TIMS data. Data bars in black are
included in the calculation of the weighed mean date. Multiple LA-ICPMS dates from the same
zircon grain are connected with thin lines.
25
population), and values <<1 suggest that the reported uncertainties are larger than what would
be expected from a single population. In detail, acceptable MSWD values actually depend on
the number of points pooled together (Wendt and Carl 1991; Spencer et al. 2016). For example,
values between 0.5 and 1.5 are acceptable for a population of 30 points (at 2σ).
However, the accuracy of weighted mean ages has been repeatedly questioned (Chiaradia et
al. 2013; Chiaradia et al. 2014; Schoene 2014). Indeed, the advent of high precision dating
techniques (CA-ID-TIMS) has highlighted that data that might look statistically equivalent at
the level of their uncertainties, can actually hide a spread of data that can only become apparent
with more precise dating methods. An illustration of this is provided in Figure 7 which shows
206
LA-ICPMS and CA-ID-TIMS Pb/238U zircon dates from a porphyry intrusion from the
Coroccohuayco porphyry-skarn deposit, Peru (Chelle-Michou et al. 2014). It is noteworthy that
those grains analyzed by CA-ID-TIMS have previously been analyzed with LA-ICPMS (with
1 to 3 spots each) before being removed from the epoxy mount for further processing. Data
points are plotted at the level of their analytical uncertainties and weighted mean dates include
additional dispersion and standard/tracer calibration uncertainties (see caption of Fig. 7 for
more details) so that they can be compared at their right level of uncertainties (i.e., neglecting
only decay constant uncertainties). Both the LA-ICPMS (36.05 ± 0.25 Ma, n = 30, MSWD =
1.3) and CA-ID-TIMS (35.639 ± 0.014 Ma, n = 7, MSWD = 1.8) weighted means yield
acceptable MSWDs (in agreement with their respective number of data points), thus suggesting
they could correspond to statistically equivalent data populations, respectively. Independently
from each other, these weighted dates would be interpreted as the age of the porphyry intrusion
at the Coroccohuayco deposit. However, Figure 7 highlights that these ages do not overlap
within uncertainties (Δt = 0.41 ± 0.25 Ma), therefore indicating that at least one of them is
inaccurate. In this case, the more precise single grain CA-ID-TIMS ages highlight more than 1
Ma of zircon crystallization in deep-seated crystal mushed (or proto-plutons) before their
incorporation into felsic melts, ascent and emplacement of the porphyry intrusion at an upper
crustal level (Chelle-Michou et al. 2014). These older zircon crystallization events cannot be
resolved at the uncertainty level of LA-ICPMS dating for which data points pool together that
are actually not part of the same population and therefore include data older than the
emplacement age, resulting in a weighted mean age that is too old. While it is common practice
in zircon CA-ID-TIMS dating to take the youngest point as best representative of the age of
magma emplacement or eruption, this practice is not appropriate for in-situ or CA-free ID-
TIMS dating techniques where the weighted mean date of the youngest cluster having an
acceptable MSWD remains the best option, although it might sometimes be slightly inaccurate.
26
This example highlights the limitations of the weighted mean approach to complex and
protracted natural processes. The statistical improvement in precision may be done at the cost
the accuracy of the dated process. The calculated weighted mean date can be either too old (e.g.,
if grains crystallized from an earlier pulse of magma are included), too young (e.g., if several
grains have suffered similar amounts of unrecognized Pb-loss) or just right by coincidence. In
fact, the time resolution of geochronology is ultimately limited by the precision of single data
points, rather than by the number of data that are pooled together to statistically reduce the age
uncertainty.
27
Figure 8. Compilation of 206Pb/238U Th-corrected dates acquired with different methods for
three porphyry intrusions at the Bajo de la Alumbrera porphyry copper deposit, Argentina. Data
are from Harris et al. (2004), Harris et al. (2008), von Quadt et al. (2011), and Buret et al.
(2016). The horizontal grey bands represent the weighted mean dates recalculated by us and
include analytical uncertainties based on U–Pb dates from tables provided in the
aforementioned publications. 1 weighted mean date reported in Harris et al. (2004). 2 weighted
mean date reported in Harris et al. (Harris et al. 2008). 3 tracer used in von Quadt et al. (2011)
(written communication to the authors). All uncertainties are given at 2σ (95% confidence).
It would be presumptuous to name the causes of these discrepancies without having the
entire set of original technical and analytical data at our disposal. Nevertheless, we can make
some conjectures. Potential causes may be: (1) that different populations of zircons grains or
domains (within a single grain) where hand-picked and dated; (2) the use of inappropriate data
reduction, common Pb correction, initial Th-correction and error propagation protocols; (3) a
distinct difference in ablation rate between sample and standard zircon resulted in inaccurate
correction for fractionation (for LA-ICPMS data); (4) inaccurate isotopic tracer calibration (for
ID-TIMS data); and/or (5) unidentified concordia parallel Pb-loss (for the LA-ICPMS data).
In the case of Bajo de la Alumbrera, the most recent data by Buret et al. (2016) are deemed
to be the most accurate (in addition of being the most precise) and tightly constrain the age of
porphyry emplacement and zircon crystallization. This example illustrates the difficulty of
dealing with legacy U–Pb data which might or might not be accurate. Obviously, there are
published ages that are inaccurate, but they would remain unnoticed until new dating is done
28
with state-of-the-art techniques. In particular, reporting of x/y/z (for ID-TIMS) and x/y/z/w (for
LA-ICPMS) uncertainties and comparison of disparate U–Pb dates at the level of their y
uncertainty should be systematic. Again, these potential biases should be carefully accounted
for when interpreting short time differences on the order of the analytical uncertainty of single
dates. This also highlight the need for thorough reporting of analytical and data handling
procedures, or even, using common analytical procedures and data reduction platforms (Košler
et al. 2013).
6. What mineral can we date with the U–Pb system and what does it date?
As of today, a great number of minerals have been used for U–Pb dating, many of which in
the context of mineral deposits. A non-exhaustive list of these minerals is provided in Table 2
which presents their main characteristics and usefulness for dating ore deposits. It is noteworthy
that this table only presents a selection of some useful minerals, but others might also be
amenable to U–Pb dating. Furthermore, ongoing and future developments will likely improve
our understanding of the U–Pb system in these and new mineral species while allowing better
precision, accuracy and interpretation of the dates.
Ideal minerals for U–Pb dating should necessarily contain traces of U, and as little common
(initial) Pb as possible. They should also have a low diffusivity for Pb so as to accurately record
the radiogenic Pb ingrowth. Many minerals used for U–Pb dating are accessory minerals
(zircon, baddeleyite, titanite, monazite, xenotime) but a handful of them are major rock forming
minerals (calcite, garnet) or even ore minerals (cassiterite, columbo-tantalite, uraninite,
wolframite) (see Table 2). This exceptional mineralogical diversity allows most types of ore
deposit and ore forming processes to be dated directly or indirectly with the U–Pb method.
However, in detail, all minerals do not provide equally precise, accurate and/or meaningful
dates. In Table 2, we have classified the minerals in three categories depending on the average
quality of the date that they can provide. Nevertheless, we stress that this classification should
only be taken as a ‘rule of thumb’ and that each case would be different. For example, zircon
might give very imprecise and discordant dates while xenotime from the same sample would
return more precise and concordant dates (e.g., Cabral and Zeh 2015).
29
Table 2. Minerals suitable for U–Pb dating in the context of mineral deposits.
Mineral Formula Main types of mineral deposit where U-Pb Event dated Principal limitations Additional comments Average Some references with application to mineral
dating can be done (non-exhaustive list) quality of deposits
U–Pb
†
dating
Allanite (Ca,Ce,La,Y)2 (Fe2+,Fe3+,Al) Skarn; IOCG: Fault-related U(±REE) Hydrothermal activity, Low to moderate amount of Possibility of Th–Pb XX Pal et al. (2011); Chen and Zhou (2014); Deng et al.
metasomatism; common Pb; No matrix-matched dating (2014)
3 O(SiO7 )(SiO4 )(OH)
magmatism standard available; High amount of
206 230
excess Pb (initial Th excess)§ ;
Apatite Ca5 (PO4 )3 (F,OH,Cl) IOCG; REE(±U,P) vein; Magmatic Ni-Cu- Cooling; hydrothermal Low to high amount of common No metamictization; X Romer (1996); Amelin et al. (1999); Gelcich et al.
PGE(±Co) sulphide; activity Pb; low U concentration; No Possibility of Th–Pb (2005); Stosch et al. (2010); Seo et al. (2015); Huston
matrix-matched international dating et al. (2016)
standard available (zircon or in-
house standard; see Chew et al.,
2011); Can be sensitive to initial
230 §
Th excess .
Baddeleyite ZrO2 Magmatic Ni-Cu-PGE(±Co) sulphide; Banded Alkaline and mafic to Crystal orientation affects Limited common Pb, no XXX Corfu and Lightfoot (1996); Schärer et al. (1997);
206
iron formation; Orogenic Au; Diamond- ultramafic magmatism, Pb/238 U ratios and dates metamictization Amelin et al. (1999); Wingate and Compston (2000);
bearing kimberlite; Rare-metal carbonatite hydrothermal activity measured with SIMS (Wingate and Müller et al. (2005); Li et al. (2005); Wu et al.
Compston, 2000); (2011); Zhang et al. (2013); Bjärnborg et al. (2015);
Wall and Scoates (2016)
Brannerite (U,Ca,Ce)(Ti,Fe)2 O6 Fault-related U(±REE-F-Ba-Th); Magmatic- Hydrothermal activity Moderate to high amount of X Frei (1996); Zartman and Smith (2009); Oberthür et
hydrothermal/epithermal U(± Ni-Co-As-Mo- common Pb; easy resetting of the al. (2009); Bergen and Fayek (2012)
Pb-PGE-Au); Archean Au paleoplacer U–Pb system (Pb loss) with
hydrothermal fluids; no matrix-
matched standard available
Calcite CaCO3 MVT Pb-Zn±F Hydrothermal activity, Moderate to high amount of Date sometimes determined X DeWolf and Halliday (1991); Brannon et al. (1996);
diagenesis common Pb; easy resetting of the with the isochron method; Coveney et al. (2000); Grandia et al. (2000); Rasbury
U–Pb system (Pb and U mobility) Inverse discordance is not and Cole (2009); Burisch et al. (2017)
with hydrothermal fluids; difficulty uncommon (U loss).
to interpret the event being dated; Mostly Pbc uncorrected
No international matrix-matched 238
U/206 Pb dates
standard available (in-house
standard)
Cassiterite SnO2 Granite-related Sn(±Mo-W-Cu-Pb-Zn-Sb-Ag) Hydrothermal activity, High amount of common Pb; No Date often determined with X Gulson and Jones (1992); Yuan et al. (2011); Chen et
greisen, skarn and lode; Supergene Sn supergene alteration (?) international matrix-matched the isochron method al. (2014); Zhang et al. (2014); Li et al. (2016)
standard available (in-house
standard)
Colombo- (Mn,Fe2+)(Nb,Ta)2 O6 Rare-metal (±Sn-W) pegmatite, greisen and Late magmatic stage, Low to moderate amount of Inverse discordance is not XXX Romer and Wright (1992); Romer and Smeds (1994);
tantalite granite hydrothermal resetting common Pb; in-situ dating often uncommon (maybe related Romer and Smeds (1996); Romer et al. (1996);
standardized to zircon mineral, the to inclusions); possible Romer and Smeds (1997); Glodny et al. (1998);
use of Coltan-139 standard is inclusions of uraninite; Smith et al. (2004); Baumgartner et al. (2006);
suggested by Che et al. (2015); can chemical abrasion is Dewaele et al. (2011); Melleton et al. (2012); Melcher
be highly metamict possible et al. (2015); Che et al. (2015); Van Lichtervelde et
al. (2016)
Garnet (Ca,Ce,La,Y)2 (Fe2+,Fe3+,Al) Skarn; Metamorphosed deposit Metasomatim, Moderate to high amount of Andradite garnet tend to X Mezger et al. (1989); Mueller et al. (1996); Glodny et
metamorphism common Pb; low U content; no have higher U content. al. (1998); Jung and Mezger (2003); Seman et al.
3 O(SiO7 )(SiO4 )(OH)
matrix-matched standard available Date sometimes determined (2017)
with the isochron method
Perovskite CaTiO3 Diamond-bearing kimberlite; Rare-metal Alkaline and ultramafic Moderate to high amount of Possibility of Th–Pb X Smith et al. (1989); Heaman (2003); Lehmann et al.
carbonatite magmatism common Pb; prone to Pb loss; in- dating (2010); Donnelly et al. (2012); Zhang et al. (2013);
situ dating often standardized to Rao et al. (2013); Wu et al. (2013a); Wu et al.
zircon mineral, perovskite standard (2013b); Griffin et al. (2014); Heaman et al. (2015);
described in Heaman (2009) Castillo-Oliver et al. (2016)
REE-Phosphate (Ce,La,Th)PO4 Rare-metal (±Sn-W) pegmatite, greisen and Hydrothermal activity, 206 Limited common Pb. No XXX Glodny et al. (1998); Torrealday et al. (2000);
High amount of excess Pb
(Monazite and granite; Orogenic Au; Banded iron formation; metamorphim, (initial
230 §
Th excess) ; strong metamictization. Petersson et al. (2001); Pigois et al. (2003); Tallarico
Xenotime) Archean Au paleoplacer; Stratabound magmatism matrix effect due to trace elements Possibility of Th–Pb et al. (2004); Salier et al. (2004); Fletcher et al.
polymetallic needs to be taken in consideration dating (2004); Schaltegger et al. (2005); Salier et al. (2005);
(Co,Cu,Pb,Zn,Fe,Au,Ag,Bi,W,REE); for SIMS dating (e.g., Fletcher et Vallini et al. (2006); Michael Meyer et al. (2006);
Unconformity-related U; MVT Pb-Zn; IOCG; al. 2010) Rasmussen et al. (2007b); Rasmussen et al. (2007a);
granite-related U-Mo; Cordilleran polymetallic Lobato et al. (2007); Mueller et al. (2007); Kempe et
206 230 al. (2008); Rasmussen et al. (2008); Vielreicher et al.
YPO4 Moderate excess Pb (intial Th
§
excess) ; strong matrix effect due to (2010); Fletcher et al. (2010); Sarma et al. (2011);
trace elements needs to be taken in Muhling et al. (2012); Aleinikoff et al. (2012a);
consideration for SIMS dating (e.g., Aleinikoff et al. (2012b); Mosoh Bambi et al. (2013);
Fletcher et al. 2010) Moreto et al. (2014); Cabral and Zeh (2015); Zi et al.
(2015); Vielreicher et al. (2015); McKinney et al.
(2015); Taylor et al. (2015); Catchpole et al. (2015);
Huston et al. (2016); Van Lichtervelde et al. (2016)
Rutile TiO2 Metamorphic and magmatic Ti; Porphyry Cu- Cooling; hydrothermal Low U concentration in most cases, XX de Ronde et al. (1992); Norcross et al. (2000); von
Au; Orogenic Au activity but rutiles from high-grade Quadt et al. (2005); Kouzmanov et al. (2009);
metamorphic rocks tend to have Morisset et al. (2009); Shi et al. (2012)
higher U contents (Meinhold 2010);
moderate amount of common Pb
Titanite CaTiOSiO4 Skarn; IOCG; Orogenic Au; VMS Late magmatic stage, Low to moderate amount of Possibility of Th–Pb XX Corfu and Muir (1989); Romer and Öhlander (1994);
(Sphene) hydrothermal activity; common Pb; Titanites BLR-1 and dating Romer et al. (1994); Eichhorn et al. (1995); Mueller
metasomatism; MKED-1 proposed as matrix- et al. (1996); Norcross et al. (2000); Salier et al.
metamorphism; (cooling) matched standards (Aleinikoff et al., (2004); Bucci et al. (2004); Wanhainen et al. (2005);
2007; Spandler et al., 2016), De Haller et al. (2006); Skirrow et al. (2007); Mueller
common use of zircon or in-house et al. (2007); Chiaradia et al. (2008); Smith et al.
standards. (2009); Li et al. (2010); Dziggel et al. (2010); Chelle-
Michou et al. (2015); Deng et al. (2015b); Deng et al.
(2015a); Seo et al. (2015); Fu et al. (2016); Poletti et
al. (2016)
Uraninite UO2 U(±Au,REE,…) deposits; Stratiform Hydrothermal activity; Prone to U and Pb mobility (loss Chemical dating (EMP) is X Hofmann and Eikenberg (1991); Hofmann and Knill
polymetallic (Co,As,Pb,Zn,Cu,Mo,…) Metamorphism and gain) and recrystallization; low common (1996); Fayek et al. (2000); Polito et al. (2005);
epithermal to high amount of common Pb; no Alexandre et al. (2007); Ono and Fayek (2011); Carl
international matrix-matched et al. (2011); Philippe et al. (2011); Dieng et al.
standard available (in-house (2013); Decree et al. (2014); Luo et al. (2015);
standard) Skirrow et al. (2016)
Zircon ZrSiO4 All type of intermediate to acidic magmatism, Magmatic, hydrothermal, Can present very complex zoning Limited common Pb. XXX Most of the references provided for the other minerals
differentiate products of mafic to ultramafic detrital provenance and patterns with several age domains Chemical abrasion is also present zircon U-Pb dating.
magmatism, Porphyry-systems (incl. deposition, metamorphic possible
Epithermal, Skarn), VMS, …
Wolframite (Fe2+,Mn)WO4 Granite-related W±Sn±Mo deposits, greisen, Hydrothermal activity Low to moderate amount of X Romer and Lüders (2005); Pfaff et al. (2009);
skarn, lodes and pegmatite common Pb; no matrix-matched Lecumberri-Sanchez et al. (2014); Harlaux et al.
standard available; Possible (2017)
alteration of the mineral; Prone to
host fluid and mineral inclusions
Other minerals: (urano)thorite, vesuvianite, U deposits; REE deposits; … Hydrothermal activity; – – – Romer (1992); Romer (1996); Rasmussen et al.
bastnaesite, polycrase, coffinite, … Metamorphism; (2008); Dill et al. (2010); Wu et al. (2011); Bergen
Supergene alteration and Fayek (2012); Dill et al. (2013); Cottle (2014);
Downes et al. (2016)
†
XXX: Low common Pb, high U and structurally robust minerals; XX: Moderate common Pb, low U and structurally robust minerals; X: Common Pb-rich, low U, structurally and/or chemically weak minerals. This classification should only be taken as a 'rule
of thumb' as each case would be different.
§ 230
Refer to Figures 4 to assess the magnitude of initial Th-disequilibrium of the expected age of the mineral.
30
6.1. Low common Pb, high U and structurally robust minerals
The most dated mineral is arguably zircon. This is mainly due to its virtual ubiquity in the
geological environment, its chemical and mechanical resistance in a range of extreme
geological processes from the surface to the deep Earth crust and to the low diffusivity of U
and Pb in its crystal lattice (Cherniak and Watson 2001; Cherniak and Watson 2003; Cherniak
et al. 1997; Harley and Kelly 2007). Importantly, zircon may contain tens to thousands of ppm
of U (Hoskin and Schaltegger 2003) while essentially excluding initial Pb upon crystallization
(Watson et al. 1997). This is mainly due to the large charge and ionic radius differences between
Pb2+ (1.26 Å) and Zr4+ (0.84 Å) in eight-fold coordination in zircon. In fact, common Pb in
zircon is often limited to small inclusions and to structurally damaged parts of the crystal which
are readily removed with a chemical abrasion procedure while preserving the crystalline portion
of the mineral (Mattinson 2005). The quality and ubiquity of this mineral has triggered most of
the technical development of U–Pb geochronology including a wealth of international reference
materials used for in-situ dating methods in all laboratories around the world.
Nevertheless, other minerals such as baddeleyite, columbite group minerals (columbo-
tantalite), and REE-phosphates (monazite and xenotime) present U enrichment and common
Pb exclusion properties comparable to zircon. Despite their occurrence in the geological
environment being more restricted than that of zircon, published data often show the same level
of precision as for zircon, according to the analytical method used. Chemical abrasion
techniques have been tested on these minerals but show contrasting behavior. In the case of
monazite and baddeleyite, chemical abrasion has not shown any significant improvement in
term of precision, reproducibility and concordance (Rioux et al. 2010; Peterman et al. 2012).
This might be due to the fact that monazite and baddeleyite do not suffer metamictization
(Seydoux-Guillaume et al. 2002, 2004; Trachenko, 2004). However, baddeleyite is suggested
to become tetragonal at high ion radiation doses, a phase change that may facilitate radiogenic
Pb mobility (Schaltegger and Davies 2017). Additionally, chemical abrasion has been
successfully applied to columbo-tantalite minerals and improved the concordance of the data
(Romer and Wright 1992). It is thought to remove small inclusions of Pb bearing minerals such
as uraninite or secondary Nb- and Ta-bearing minerals (Romer et al. 1996).
31
sufficiently enriched in U to allow precise dating in most cases. Analytical protocols and
matrix-matched standards for in-situ dating have been developed and allow some labs to
routinely date these mineral (Storey et al. 2006; Aleinikoff et al. 2007; Storey et al. 2007;
Gregory et al. 2007; Luvizotto et al. 2009; Zack et al. 2011; Darling et al. 2012; Schmitt and
Zack 2012; Smye et al. 2014). The use of titanite and especially rutile as geochronometers
might be limited by their relatively lower closure temperature of the U–Pb system compared to
zircon. Hydrothermal titanite (e.g., in skarn deposits) would crystallize near or just below its
closure temperature allowing its use as a geochronometer (Chiaradia et al. 2008; Chelle-Michou
et al. 2015), and helping to pinpoint antecrystic zircon growth (i.e., crystallized in earlier
magma pulses and incorporated in a later pulse; Miller et al. 2007) in the skarn-forming
magmatic intrusion. Rutile is involved in high temperature metamorphic reactions and can
produce new zircon upon recrystallization at lower temperature and expulsion of Zr (e.g., Pape
et al. 2016). Allanite may have exceedingly high Th/U ratios requiring a very careful approach
for accurately correcting and interpreting initial 230Th disequilibrium (Oberli et al. 2004).
32
Due to the ubiquity of calcite in vein, cement or replacement phase in mineral deposits, calcite
U–Pb dating is expected to open to new opportunities for ore deposit research and to address
the timing of crustal fluid flow through direct dating. Yet, the main difficulty of calcite dating
is to correctly interpret the event being dated, or if unsure, allow for all reasonable possibilities
(e.g., see the case of the Hamersley spherule beds, Australia; Woodhead et al. 1998; Rasmussen
et al. 2005).
33
Despite its common usage in ore deposit research, the dating of magma intrusion only rarely
dates the mineralization itself. In fact, this is only restricted to places where the ore minerals
have crystallized under magmatic conditions such as the magmatic Ni-Cu-Cr(±Au±PGE)
deposits and possibly some magmatic REE deposits as well. If appropriate crosscutting
relationship with the mineralization can be observed, dating magmatic intrusions can elegantly
bracket the timing of ore deposition (e.g., von Quadt et al. 2011). In the case of porphyry,
greisen, or VMS deposits the age of the ore-related intrusion or of the associated volcanics may
often provide a good, if not excellent, approximation for age of the mineralization. Yet, this
approach requires much caution as even in classical magmatic-hydrothermal deposits such as
W-Sn granite deposits or porphyry Cu deposits, the mineralization can have been sourced by a
hidden intrusion at depth while being hosted in a previously emplaced one (e.g., Schaltegger et
al. 2005). However, for deposits where the relationship between ore formation and a particular
magma intrusion is ambiguous (e.g., IOCGs, orogenic Au deposit, epithermal deposits, distal
skarns) or even totally absent (e.g., MVT deposits) it is much more advantageous to determine
directly the timing of hydrothermal fluid circulation and/or of ore deposition. The list of ore
minerals suitable for U–Pb dating include cassiterite (for Sn deposits), wolframite (for W
deposits) columbo-tantalite (in some rare-metal granite, greisen and pegmatite deposits), rutile
(for Ti deposits), and minerals associated with U deposits (e.g., uraninite, brannerite). This
restricts the types of ore that can dated with the U–Pb method. Alternatively, several gangue
mineral species can be used to date hydrothermal fluid circulation, metasomatism and
metamorphism. Their relevance for the genesis or reworking of the studied ore deposit is
fundamentally linked to their position in the paragenetic sequence with respect to the ore
minerals. REE-phosphates such as monazite and xenotime are common in a wide variety of
hydrothermal systems ranging from granite-related rare metal deposits to MVT deposits (Table
2) and, if available, would be the ideal minerals to date hydrothermal processes. In few cases,
hydrothermal zircons at skarn (Niiranen et al. 2007; Wan et al. 2012; Deng et al. 2015c), IOCG
(Valley et al. 2009), orogenic Au (Kerrich and Kyser 1994; Pelleter et al. 2007), and
alkaline/carbonatite magmatism related rare-metal deposits (Yang et al. 2013; Campbell et al.
2014) have been reported and can date hydrothermal activity and metasomatism. However, in
the absence of these hydrothermal minerals (which is not uncommon), other minerals listed in
Table 2 with non-negligible amounts of common Pb can be called on. Titanite or allanite can
provide excellent dates for skarn (Chiaradia et al. 2008; Deng et al. 2014; Chelle-Michou et al.
2015; Deng et al. 2015b) and IOCG deposits (Skirrow et al. 2007; Smith et al. 2009; Haller et
al. 2011). Ore-stage calcite or apatite may sometimes represent the only minerals suitable for
34
U–Pb dating at MVT deposits (Grandia et al. 2000) or some REE-P deposits (Huston et al.
2016). The ability of apatite to keep record of Cl, F, OH and SO42- of the hydrothermal fluid
(or magma) from which it crystallizes (Webster and Piccoli 2015; Harlov 2015) coupled with
the possibility to date it with the U–Pb method (Chew and Spikings 2015) opens interesting
opportunities to refine ore forming models. Finally, U–Pb minerals such as rutile, apatite and/or
titanite can provide invaluable thermochronological information on the thermal evolution of
the studied ore deposit during and after its genesis.
7.1. Input of multi-mineral U–Pb dating for understanding gold deposition and remobilization
in the Witwatersrand basin, South Africa
About 32 % of all gold ever mined and about the same proportion of known gold resources
comes from deposits hosted in the Witwatersrand Basin, South Africa (Frimmel and Hennigh
2015), a Mesoarchean detrital sedimentary basin deposited on the Kaapvaal Craton (Fig. 9).
The genesis of this enormous accumulation of gold in the crust has triggered one of the “greatest
debate in the history of economic geology” (see summary in Muntean et al. 2005). Proposed
models for the deposition of gold range from a modified paleoplacer to a purely hydrothermal
origin. These disparate views arises from contradicting observations that are selectively put
forward to favor either model (Frimmel et al. 2005; Law and Phillips 2005; Muntean et al.
2005). In fact, probably none of these end-member models can account for all the geological,
chemical and isotopic observations. The most recent models rather consider the very peculiar
conditions that prevailed in the Mesoarchean atmosphere, hydrosphere and biosphere (Frimmel
and Hennigh 2015; Heinrich 2015). At this time, redox exchanges mediated by microbial life
35
could have triggered the synsedimentary precipitation of the large quantities of gold dissolved
in acidic and reduced meteoric and shallow sea waters.
Figure 9. Compilation of available U–Pb data from the Witwatersrand basin plotted against the
stratigraphic position of the sample. Stratigraphic column of the Archean to early Proterozoic
succession in South Africa from Muntean et al. (2005). Data from the Witwatersrand basin are
from Armstrong et al. (1991), England et al. (2001), Kositcin et al. (2003) Kositcin and Krapež
(2004), Rasmussen et al. (2007a) and Koglin et al. (2010). Ages of the intrusion of the Bushveld
Complex and of the Vredefort impact are from Zeh et al. (2015) and Moser (1997), respectively.
Kernel density estimates (KDE) where obtained using DensityPlotter (Vermeesch 2012).
Selected data are 95-105 % concordant and data points are plotted at the 2σ uncertainty level.
Inset map of the Kaapvaal craton is modified from Poujol (2007).
36
the sediments (Fig. 9). One of the most significant contribution comes from Armstrong et al.
(1991) who dated zircons from volcanic rocks distributed along the sedimentary pile of the
basin. They constrained the deposition of the Witwatersrand Supergroup to within a timeframe
of ca. 360 Ma from 3074 to 2714 Ma. Subsequent studies have focused on detrital zircon and
xenotime from the main formations present along the stratigraphic column and intimately
associated with the gold-bearing reefs (England et al. 2001; Kositcin and Krapež 2004; Koglin
et al. 2010). These have confirmed the previous depositional ages but provide additional insight
in the source of the detritus that filled the basin, as well as secular changes in the catchment
area of the basin over time. Results show that the source area of detritus has an increasing age-
range of rocks undergoing erosion over time. Apart from the lowermost part of the
Witwatersrand Supergroup (Orange Formation, West Rand Group) which has dates clustering
around 3.21 Ga, zircon dates from the West Rand Group cluster around 3.06 Ga, with only few
older and younger dates (Fig. 9). Furthermore, zircon dates from the Central Rand Group shows
additional peaks at 2.96-2.92 Ga and 3.44-3.43 Ga with several intervening dates in between
these main peaks. Dates of detrital xenotime are mostly within the 3.1-2.9 Ga range but also
extend as low as 2.8 Ga (Fig. 9). Koglin et al. (2010) and Ruiz et al. (2006) further link the
gold-rich sediments to the presence of the 3.06 Ga zircon age peak. When compared with
outcropping Archean terrains of the Kaapvaal Craton, these zircons could have originated from
the greenstone belts west of the Witwatersrand basin (Madibe and Kraaipan), rocks in the
immediate proximity of the basin (e.g., Johannesburg and Vrefefort Dome) or from equivalent
units located northwest of the basin that might be present below the post-Witwatersrand cover
(Koglin et al. 2010). More distal candidates such as the Murchison and the Barberton
greenstone belts have also been proposed (Ruiz et al. 2006; Koglin et al. 2010). This
interpretation is also compatible with paleocurrent directions and isotopic data (Koglin et al.
2010).
A paleoplacer model requires that all of the gold deposited in the basin originated from the
same eroding massifs that sourced the sediments. However, such gigantic quantities of gold are
two orders of magnitude in excess of all the gold ever mined and discovered in the potential
outcropping massifs that sourced the zircons. This observation has been a major argument
against any sort of paleoplacer model (e.g., Phillips and Law 2000; Law and Phillips 2005;
Frimmel and Hennigh 2015). The existence of a now vanished or buried, hypothetical massif
as a source of this huge amount of gold would pose an equally important question about how
this massif would have been exceptionally well endowed with gold.
37
An epigenetic (hydrothermal) origin of the gold is supported by several petrographic
observations. Yet, cross-cutting relationship suggest that hydrothermal activity took place
before deposition of the Platberg Group, that is, before ca. 2.7 Ga (e.g., Law and Phillips 2005;
Meier et al. 2009). U–Pb dating of diagenetic xenotime have yielded a major peak between
2.78-2.72 Ga which could be related to a heating event and flood-basalt volcanism during the
deposition of the Klipriviersberg Group, immediately following the deposition of the
Witwatersrand Supergroup (Fig. 9; England et al. 2001; Kositcin et al. 2003). Although this
timing for gold introduction would be consistent with temporal constrains, the association of
gold with this 2.78-2.72 Ga xenotime has not been reported. Additionally, U–Pb dating of
metamorphic-hydrothermal REE-phosphates (monazite and xenotime) paragenetically
associated with some gold or unrelated to gold mostly records ages between 2.06-2.03 Ga
throughout the stratigraphic succession from the Witwatersrand to the Transvaal Supergroups
(Fig. 9; England et al. 2001; Kositcin et al. 2003; Rasmussen et al. 2007a). This age is consistent
with the emplacement age 1of the Bushveld complex on the northern flank of the Witwatersrand
Basin (Zeh et al. 2015) which most likely triggered fluid circulation, gold remobilization and
peak greenschist metamorphic conditions in the basin (Rasmussen et al. 2007a).
While none of the available U–Pb data for the Witwatersrand basin (Fig. 9) can firmly date
gold deposition, or conclusively explain how gold was deposited, they have provided the
necessary temporal framework on which to challenge relative chronological data. They have
brought significant arguments against each of the classical models invoked for the formation of
this district (syngenetic vs epigenetic) while confirming that gold remobilization occurred long
after the formation of the deposit and contributed to the emergence of new ore forming models
(Frimmel and Hennigh 2015; Heinrich 2015). This example highlights the necessity to properly
constrain each U–Pb date against paragenetic, cross-cutting and stratigraphic observations in
order to draw meaningful conclusions. The Witwatersrand gold deposits result from a long-
lived and multi-episodic geological history where U–Pb geochronology provided constraints
on basin formation, sediment provenance, diagenesis and metamorphism. It is noteworthy that
the different minerals that were dated (zircon, monazite, xenotime), individually record a
limited portion of the multiple processes that shaped the Witwatersrand basin and proved to be
highly complementary to each other. Unveiling this protracted history did not require
particularly high-precision dating methods, as LA-ICPMS and SIMS instruments with high
sample throughput (Table 1) proved very affective. Additionally, the very high spatial
resolution achievable with a SIMS instrument was crucial in unlocking the U–Pb information
in tiny xenotime and monazite crystals identified from thin sections.
38
7.2. Zircon U–Pb insights into the genesis on porphyry copper deposits
Porphyry copper deposits (PCDs) typically form at convergent margins in association with
subduction or post-subduction magmatism (e.g., Richards 2009). Metals and sulfur fixed in
these deposits are thought to have been sourced from a cooling and degassing fluid-saturated
magma body emplaced at shallow depths within the upper crust and transported to the site of
deposition by magmatic-hydrothermal fluids (Hedenquist and Lowenstern 1994; Sillitoe 2010;
Pettke et al. 2010; Simon and Ripley 2011; Richards 2011). Ultimately, very efficient fluid
focusing and sulfide precipitation together with post-mineralization ore deposit preservation
will favor the presence of economic porphyry deposits at erosion levels.
Figure 10. (A) probability density distribution of Cu endowment in global porphyry copper
deposits (PCDs). Data from Singer et al. (2008). (B) Correlation between the duration of the
mineralizing event and the total amount of Cu deposited (adapted from Chiaradia and Caricchi,
2017). BH: Batu Hijau (Indonesia), BjA: Bajo de la Alumbrera (Argentina), Co:
Coroccohuayco (Peru), EA: El Abra (Chile), BB: Boyongan-Bayugo (Philippines)), Bh:
Bingham (US), Cha: Chaucha (Ecuador), Ju: Junin (Ecuador), ES: El Salvador (Chile), Es:
Escondida (Chile), LP: Los Pelambres (Chile), Chu: Chuquicamata (Chile), RB: Rio Blanco
(Chile), Bt: Butte (US), ET: El Teniente (Chile), RD: Reko Diq (Pakistan).
39
The USGS global database of PCDs shows that these deposits span more than four orders of
magnitude in copper endowment (Singer et al. 2008; Fig. 10a). Yet, the specific factors that
control the size of these deposits have remained speculative. Comparing ‘standard’ and ‘giant’
PCDs, Richards (2013) speculated that the formation of the largest deposits result from a
combination of copper enrichment in the magma, the focusing of fluids in structural corridors
and, long-lived hydrothermal activity may favor the formation of the largest deposits. Among
these possible factors, the timescale of PCD formation may play a significant role in their size.
Compiling geochronological data (U–Pb on zircon and Re-Os on molybdenite) from PCDs
around the world, Chelle-Michou et al. (2017) and Chiaradia and Caricchi (2017) have
highlighted a correlation between the duration of the mineralizing event and the total mass of
copper deposited, suggesting an average copper deposition rate of about 40 Mt/Ma (Fig. 10b).
This relationship probably reflects the mass balance requirement for a giant deposit to be
sourced by a large body of magma, which is incrementally injected into the upper crust over
long timescales (see Chelle-Michou et al. 2017).
Similar conclusions where reached by Caricchi et al. (2014) who suggested that magmatism
associated with economic PCDs is distinguishable from background pluton-forming
magmatism and large-eruption-forming magmatism by large magma volumes emplaced at
average rate of magma injection (~0.001 km3/yr). This conclusion was drawn through inverse
thermal modelling of high-precision CA-ID-TIMS U–Pb zircon age distributions.
While geochronology on PCDs has been mostly used to determine the formation age of these
deposits, high-precision geochronological data can now be used to elucidate the duration of the
ore-forming process. Figure 10b shows that the duration of ore-formation may be a significant
control on their size (i.e., metal endowment) and, by inference, on the specific processes
responsible for their formation. In addition, high-precision geochronological data may be able
to help test the validity of numerical models of PCD formation (e.g., Weiss et al., 2012), or
directly as input data into numerical models aiming at quantifying the time-volume-flux-
geochemistry relationships of the magmatism associated with PCD genesis (e.g., Caricchi et al.
2014; Chelle-Michou et al. 2017; Chiaradia and Caricchi 2017). These studies only start to
unearth the great potential of high-precision CA-ID-TIMS U–Pb dating for PCD exploration,
and can also significantly contribute to a better understanding of PCD magmatic ore-forming
processes.
Concluding remarks
40
Over the past two decades U–Pb geochronology has become an essential tool for the study
of ore deposits. After a century of development, more than 16 minerals can now be dated with
the U–Pb technique allowing its use for most types of ore deposits. U–Pb dating is most
commonly used to provide the age of a particular geological event related to a studied deposit
(e.g., magmatism, hydrothermal activity, sedimentation, metamorphism, ore deposition and
remobilization), depending on the mineral(s) available for dating. The choice of the mineral(s)
and of the analytical technique (LA-ICPMS, SIMS or ID-TIMS) used for dating mainly
depends on the scientific questions that need to be answered and on the opportunities offered
by the studied deposit. This point is perhaps one of the main limitations of the U–Pb dating of
ore deposits. For example, MVT deposits rarely contain minerals suitable for U–Pb dating
(potentially calcite, provided it has low initial Pb), in which case the use of other isotopic
systems will be necessary (e.g., Rb-Sr on sphalerite, Re-Os on sulfides). In addition, as we have
seen in the case of the Witwatersrand basin, the spatial resolution required for the analysis may
sometimes critical guide the choice of the analytical method.
Recent advances that combine numerical modelling with U–Pb geochronology for porphyry
copper deposits suggest that high-precision zircon U–Pb data may also be used as a window to
better understand the magmatic aspect of the ore-forming process (Caricchi et al. 2014) and to
unravel the fundamental controls on the size of the deposit (Chelle-Michou et al. 2017;
Chiaradia and Caricchi, 2017). Comparable studies on other deposit types could potentially
advance our understanding of ore-forming processes and may generate innovative tools for
mineral exploration.
A further important development of U–Pb geochronology concerns its coupling with textural
and geochemical data (e.g., trace elements, Lu–Hf isotopes, O isotopes) obtained on the same
grain or on the same spot as the U–Pb data. This is commonly referred to as ‘petrochronology’
and allows temporal information relative to the evolution and/or the source of the liquid (a
magma or an aqueous fluid) from which the mineral precipitated and potentially the rate of its
evolution (e.g., Ballard et al. 2002; Smith et al. 2009; Valley et al. 2010; Pal et al. 2011; Rao et
al. 2013; Yang et al. 2013; Griffin et al. 2014; Rezeau et al. 2016; Poletti et al. 2016; Gardiner
et al. 2017). In particular, high-precision petrochronology on zircon and baddeleyite can
provide unprecedented insight into the processes at play during magma evolution, the potential
turning point leading to mineralization, or, the link between small intrusive bodies (dykes or
stocks) or volcanic products and their larger deep-seated plutonic source (e.g., Wotzlaw et al.
2013; Chelle-Michou et al. 2014; Wotzlaw et al. 2015; Tapster et al. 2016; Buret et al. 2016;
2017; Schaltegger and Davies, 2017).
41
The field of U–Pb geochronology is working towards a level of maturity whereby inter-
laboratory reproducibility will be guaranteed in most labs around the world and where each
date and its uncertainty can be fully traceable to SI units. This however, should not mask the
high-level of competency and training required to certify the quality of the analysis, to maintain
the lab at the best level (picogram levels of common Pb contaminations can be dramatic in a
CA-ID-TIMS lab) and, very importantly, the interpretation of the dates into geologically
relevant ages. As we have shown, there are numerous potential pitfalls that, if not carefully
accounted for, can result in unsupported or even wrong conclusions.
The improving precision, accuracy and spatial resolution of analyses now achievable,
challenges paradigms of ore-forming processes and will continue to contribute significant
breakthroughs in ore deposit research and potentially also contribute to the development of new
mineral exploration tools. The full added value of U–Pb geochronology will however only be
assured through its coupling with geochemical data, high-quality field and petrographic
observations and numerical modelling.
Acknowledgments
We are grateful to the D. Huston for inviting us to write this chapter. Constructive reviews from
Andrew Cross and Neal McNaughton significantly improved the clarity of the manuscript and
were very much appreciated. C. Chelle-Michou acknowledges financial support from the
European Commission and the Université Jean Monnet during the preparation of the
manuscript. Both authors acknowledge the support of the Swiss National Science Foundation.
References
Aleinikoff JN, Hayes TS, Evans KV, Mazdab FK, Pillers RM, Fanning CM (2012a) SHRIMP
U–Pb Ages of xenotime and monazite from the Spar Lake red bed-associated Cu-Ag
deposit, western Montana: implications for ore genesis. Econ Geol 107:1251–1274. doi:
10.2113/econgeo.107.6.1251
Aleinikoff JN, Slack JF, Lund K, Evans KV, Fanning CM, Mazdab FK, Wooden JL, Pillers
RM (2012b) Constraints on the timing of Co-Cu ± Au mineralization in the Blackbird
District, Idaho, using SHRIMP U–Pb ages of monazite and xenotime plus zircon ages of
related Mesoproterozoic orthogneisses and metasedimentary rocks. Econ Geol 107:1143–
1175. doi: 10.2113/econgeo.107.6.1143
Aleinikoff JN, Wintsch RP, Tollo RP, Unruh DM, Fanning CM, Schmitz MD (2007) Ages
and origins of rocks of the Killingworth dome, south-central Connecticut: implications for
the tectonic evolution of southern New England. Am J Sci 307:63–118. doi:
10.2475/01.2007.04
42
Alexandre P, Kyser K, Thomas D, Polito P, Marlat J (2007) Geochronology of unconformity-
related uranium deposits in the Athabasca Basin, Saskatchewan, Canada and their
integration in the evolution of the basin. Miner Deposita 44:41–59. doi: 10.1007/s00126-
007-0153-3
Anczkiewicz R, Oberli F, Burg JP, Villa IM, Günther D, Meier M (2001) Timing of normal
faulting along the Indus suture in Pakistan Himalaya and a case of major 231Pa/235U initial
disequilibrium in zircon. Earth Planet Sci Lett 191:101–114. doi: 10.1016/S0012-
821X(01)00406-X
Arevalo R Jr (2014) Laser ablation ICP-MS and laser fluorination GS-MS. In: Treatise on
Geochemistry. Elsevier, pp 425–441
Armstrong RA, Compston W, Retief EA, Williams IS, Welke HJ (1991) Zircon ion
microprobe studies bearing on the age and evolution of the Witwatersrand triad.
Precambrian Res 53:243–266. doi: 10.1016/0301-9268(91)90074-K
Aston FW (1929) The mass-spectrum of uranium lead and the atomic weight of protactinium.
Nature 123:313. doi: 10.1038/123313a0
Ballard J, Palin M, Campbell I (2002) Relative oxidation states of magmas inferred from
Ce(IV)/Ce(III) in zircon: application to porphyry copper deposits of northern Chile.
Contrib Mineral Petrol 144:347–364. doi: 10.1007/s00410-002-0402-5
Ballard JR, Palin JM, Williams IS, Campbell IH, Faunes A (2001) Two ages of porphyry
intrusion resolved for the super-giant Chuquicamata copper deposit of northern Chile by
ELA-ICP-MS and SHRIMP. Geology 29:383–386. doi: 10.1130/0091-
7613(2001)029<0383:TAOPIR>2.0.CO;2
Barboni M, Annen C, Schoene B (2015) Evaluating the construction and evolution of upper
crustal magma reservoirs with coupled U/Pb zircon geochronology and thermal modeling:
a case study from the Mt. Capanne pluton (Elba, Italy). Earth Planet Sci Lett 432:436–
448. doi: 10.1016/j.epsl.2015.09.043
43
bearing granitic pegmatites from the Seridó belt, northeastern Brazil: genetic constraints
from U–Pb dating and Pb isotopes. Can Mineral 44:69–86. doi:
10.2113/gscanmin.44.1.69
Becquerel H (1896a) Sur les radiations invisibles émises par les corps phosphorescents. C R
Acad Sci Paris 122:501–503.
Becquerel H (1896b) Sur les radiations invisibles émises par les sels d'uranium. C R Acad Sci
Paris 122:689–694.
Bergen L, Fayek M (2012) Petrography and geochronology of the Pele Mountain quartz-
pebble conglomerate uranium deposit, Elliot Lake District, Canada. Am Min 97:1274–
1283. doi: 10.2138/am.2012.4040
Beyer C, Berndt J, Tappe S, Klemme S (2013) Trace element partitioning between perovskite
and kimberlite to carbonatite melt: new experimental constraints. Chem Geol 353:132–
139. doi: 10.1016/j.chemgeo.2012.03.025
Black LP, Kamo SL, Allen CM, Davis DW, Aleinikoff JN, Valley JW, Mundil R, Campbell
IH, Korsch RJ, Williams IS, Foudoulis C (2004) Improved 206Pb/238U microprobe
geochronology by the monitoring of a trace-element-related matrix effect; SHRIMP, ID–
TIMS, ELA–ICP–MS and oxygen isotope documentation for a series of zircon standards.
Chem Geol 205:115–140. doi: 10.1016/j.chemgeo.2004.01.003
Boltwood BB (1907) Ultimate disintegration products of the radioactive elements; Part II,
Disintegration products of uranium. Am J Sci Series 4 Vol. 23:78–88. doi:
10.2475/ajs.s4-23.134.78
Bowring JF, McLean NM, Bowring SA (2011) Engineering cyber infrastructure for U–Pb
geochronology: Tripoli and U-Pb_Redux. Geochem Geophys Geosyst 12:Q0AA19. doi:
10.1029/2010GC003479
Brannon JC, Cole SC, Podosek FA, Ragan VM, Coveney RM, Wallace MW, Bradley AJ
(1996) Th–Pb and U–Pb Dating of ore-stage calcite and Paleozoic fluid flow. Science
44
271:491–493. doi: 10.1126/science.271.5248.491
Bucci LA, McNaughton NJ, Fletcher IR, Groves DI, Kositcin N, Stein HJ, Hagemann SG
(2004) Timing and duration of high-temperature gold mineralization and spatially
associated granitoid magmatism at Chalice, Yilgarn Craton, Western Australia. Econ
Geol 99:1123–1144. doi: 10.2113/gsecongeo.99.6.1123
Buret Y, von Quadt A, Heinrich C, Selby D, Wälle M, Peytcheva I (2016) From a long-lived
upper-crustal magma chamber to rapid porphyry copper emplacement: reading the
geochemistry of zircon crystals at Bajo de la Alumbrera (NW Argentina). Earth Planet
Sci Lett 450:120–131. doi: 10.1016/j.epsl.2016.06.017
Buret Y, Wotzlaw J-F, Roozen S, Guillong M, von Quadt A, Heinrich CA (2017) Zircon
petrochronological evidence for a plutonic-volcanic connection in porphyry copper
deposits. Geology 45:623–626. doi: 10.1130/G38994.1
Burisch M, Gerdes A, Walter BF, Neumann U, Fettel M, Markl G (2017) Methane and the
origin of five-element veins: mineralogy, age, fluid inclusion chemistry and ore forming
processes in the Odenwald, SW Germany. Ore Geol Rev 81:42–61.
Cabral AR, Zeh A (2015) Detrital zircon without detritus: a result of 496-Ma-old fluid–rock
interaction during the gold-lode formation of Passagem, Minas Gerais, Brazil. Lithos 212-
215:415–427. doi: 10.1016/j.lithos.2014.10.011
Campbell LS, Compston W, Sircombe KN, Wilkinson CC (2014) Zircon from the East
Orebody of the Bayan Obo Fe–Nb–REE deposit, China, and SHRIMP ages for
carbonatite-related magmatism and REE mineralization events. Contrib Mineral Petrol
168:1041. doi: 10.1007/s00410-014-1041-3
Caricchi L, Simpson G, Schaltegger U (2014) Zircons reveal magma fluxes in the Earth's
crust. Nature 511:457–461. doi: 10.1038/nature13532
Carl C, Pechmann von E, Höhndorf A, Ruhrmann G (2011) Mineralogy and U/Pb, Pb/Pb, and
Sm/Nd geochronology of the Key Lake uranium deposit, Athabasca Basin,
Saskatchewan, Canada. Can J Earth Sci 29:879–895. doi: 10.1139/e92-075
Castillo-Oliver M, Galí S, Melgarejo JC, Griffin WL, Belousova E, Pearson NJ, Watangua M,
O'Reilly SY (2016) Trace-element geochemistry and U–Pb dating of perovskite in
kimberlites of the Lunda Norte province (NE Angola): Petrogenetic and tectonic
implications. Chem Geol 426:118–134. doi: 10.1016/j.chemgeo.2015.12.014
Chakhmouradian AR, Reguir EP, Kamenetsky VS, Sharygin VV, Golovin AV (2013) Trace-
element partitioning in perovskite: implications for the geochemistry of kimberlites and
other mantle-derived undersaturated rocks. Chem Geol 353:112–131. doi:
45
10.1016/j.chemgeo.2013.01.007
Che XD, Wu F-Y, Wang R-C, Gerdes A, Ji W-Q, Zhao Z-H, Yang J-H, Zhu Z-Y (2015) In
situ U–Pb isotopic dating of columbite–tantalite by LA–ICP–MS. Ore Geol Rev 65:979–
989. doi: 10.1016/j.oregeorev.2014.07.008
Chen WT, Zhou M-F (2014) Ages and compositions of primary and secondary allanite from
the Lala Fe–Cu deposit, SW China: implications for multiple episodes of hydrothermal
events. Contrib Mineral Petrol 168:1043. doi: 10.1007/s00410-014-1043-1
Chen X-C, Hu R-Z, Bi X-W, Li H-M, Lan J-B, Zhao C-H, Zhu J-J (2014) Cassiterite LA-
MC-ICP-MS U/Pb and muscovite 40Ar/39Ar dating of tin deposits in the Tengchong-
Lianghe tin district, NW Yunnan, China. Miner Deposita 49:843–860. doi:
10.1007/s00126-014-0513-8
Cheng H, Lawrence Edwards R, Shen C-C, Polyak VJ, Asmerom Y, Woodhead J, Hellstrom
J, Wang Y, Kong X, Spötl C, Wang X, Calvin Alexander E (2013) Improvements in
230
Th dating, 230Th and 234U half-life values, and U–Th isotopic measurements by multi-
collector inductively coupled plasma mass spectrometry. Earth Planet Sci Lett 371:82–91.
doi: 10.1016/j.epsl.2013.04.006
Cherniak DJ, Hanchar JM, Watson EB (1997) Diffusion of tetravalent cations in zircon.
Contrib Mineral Petrol 127:383–390. doi: 10.1007/s004100050287
Cherniak DJ, Watson EB (2001) Pb diffusion in zircon. Chem Geol 172:5–24. doi:
10.1016/S0009-2541(00)00233-3
Cherniak DJ, Watson EB (2003) Diffusion in zircon. Rev Mineral Geochem 53:113–143. doi:
10.2113/0530113
Chew DM, Spikings RA (2015) Geochronology and thermochronology using apatite: time
and temperature, lower crust to surface. Elements 11:189–194. doi:
10.2113/gselements.11.3.189
Chew DM, Sylvester PJ, Tubrett MN (2011) U–Pb and Th–Pb dating of apatite by LA-
ICPMS. Chem Geol 280:200–216. doi: 10.1016/j.chemgeo.2010.11.010
46
Chiaradia M, Caricchi L (2017) Stochastic modelling of deep magmatic controls on porphyry
copper deposit endowment. Sci Rep 7:44523. doi: 10.1038/srep44523
Clemens J (2003) S-type granitic magmas – petrogenetic issues, models and evidence. Earth
Sci Rev 61:1–18. doi: 10.1016/S0012-8252(02)00107-1
Cochrane R, Spikings RA, Chew DM, Wotzlaw J-F, Chiaradia M, Tyrrell S, Schaltegger U,
van der Lelij R (2014) High temperature (>350°C) thermochronology and mechanisms of
Pb loss in apatite. Geochim Cosmochim Acta 127:39–56. doi: 10.1016/j.gca.2013.11.028
Condon DJ, Schoene B, McLean NM, Bowring SA, Parrish RR (2015) Metrology and
traceability of U–Pb isotope dilution geochronology (EARTHTIME Tracer Calibration
Part I). Geochim Cosmochim Acta 164:464–480. doi: 10.1016/j.gca.2015.05.026
Coogan LA, Parrish RR, Roberts NMW (2016) Early hydrothermal carbon uptake by the
upper oceanic crust: Insight from in situ U-Pb dating. Geology 44:147–150. doi:
10.1130/G37212.1
Corfu F (2013) A century of U–Pb geochronology: the long quest towards concordance. Geol
Soc Am Bull 125:33–47. doi: 10.1130/B30698.1
Corfu F, Muir TL (1989) The Hemlo-Heron Bay greenstone belt and Hemlo Au-Mo deposit,
Superior Province, Ontario, Canada 2. Timing of metamorphism, alteration and Au
mineralization from titanite, rutile, and monazite U–Pb geochronology. Chem Geol
79:201–223. doi: 10.1016/0168-9622(89)90030-4
Coveney RM, Ragan VM, Brannon JC (2000) Temporal benchmarks for modeling
Phanerozoic flow of basinal brines and hydrocarbons in the southern Midcontinent based
on radiometrically dated calcite. Geology 28:795–798. doi: 10.1130/0091-
47
7613(2000)028<0795:tbfmpf>2.3.co;2
Curie P, Curie M, Bémont G (1898) Sur une nouvelle substance fortement radio-active
contenue dans la pechblende. C R Acad Sci Paris 127:1215–1217.
Curie P, Skolodowska Curie M (1898) Sur une substance nouvelle radioactive, contenue dans
la pechblende. C R Acad Sci Paris 127:175–178.
Darling JR, Storey CD, Engi M (2012) Allanite U–Th–Pb geochronology by laser ablation
ICPMS. Chem Geol 292-293:103–115.
Davis DW, Krogh TE, Williams IS (2003) Historical development of zircon geochronology.
Rev Mineral Geochem 53:145–181. doi: 10.2113/0530145
de Ronde CEJ, Spooner ETC, de Wit MJ, Bray CJ (1992) Shear zone-related, Au quartz vein
deposits in the Barberton greenstone belt, South Africa; field and petrographic
characteristics, fluid properties, and light stable isotope geochemistry. Econ Geol 87:366–
402. doi: 10.2113/gsecongeo.87.2.366
Decree S, Deloule É, De Putter T, Dewaele S, Mees F, Baele J-M, Marignac C (2014) Dating
of U-rich heterogenite: new insights into U deposit genesis and U cycling in the Katanga
Copperbelt. Precambrian Res 241:17–28.
Deng X-D, Li J-W, Wen G (2014) Dating iron skarn mineralization using hydrothermal
allanite-(La) U–Th–Pb isotopes by laser ablation ICP-MS. Chem Geol 382:95–110.
Deng X-D, Li J-W, Zhao X-F, Wang H-Q, Qi L (2015a) Re–Os and U–Pb geochronology of
the Laochang Pb–Zn–Ag and concealed porphyry Mo mineralization along the
Changning–Menglian suture, SW China: implications for ore genesis and porphyry Cu–
Mo exploration. Miner Deposita 51:237–248. doi: 10.1007/s00126-015-0606-z
Deng X-D, Li J-W, Zhou M-F, Zhao X-F, Yan D-R (2015b) In-situ LA-ICPMS trace
elements and U–Pb analysis of titanite from the Mesozoic Ruanjiawan W–Cu–Mo skarn
deposit, Daye district, China. Ore Geol Rev 65:990–1004. doi:
10.1016/j.oregeorev.2014.08.011
48
Deng XD, Li JW, Wen G (2015c) U–Pb geochronology of hydrothermal zircons from the
Early Cretaceous iron skarn deposits in the Handan-Xingtai district, North China Craton.
Econ Geol 110:2159–2180. doi: 10.2113/econgeo.110.8.2159
Dieng S, Kyser K, Godin L (2013) Tectonic history of the North American shield recorded in
uranium deposits in the Beaverlodge area, northern Saskatchewan, Canada. Precambrian
Res 224:316–340.
Dill HG, Gerdes A, Weber B (2010) Age and mineralogy of supergene uranium minerals —
Tools to unravel geomorphological and palaeohydrological processes in granitic terrains
(Bohemian Massif, SE Germany). Geomorphology 117:44–65. doi:
10.1016/j.geomorph.2009.11.005
Dill HG, Hansen BT, Weber B (2013) U/Pb age and origin of supergene uranophane-beta
from the Borborema Pegmatite mineral province, Brazil. J South Am Earth Sci 45:160–
165. doi: 10.1016/j.jsames.2013.03.014
Donnelly CL, Griffin WL, Yang J-H, O'Reilly SY, Li Q-L, Pearson NJ, Li X-H (2012) In situ
U–Pb dating and Sr–Nd isotopic analysis of perovskite: constraints on the age and
petrogenesis of the Kuruman Kimberlite province, Kaapvaal Craton, South Africa. J
Petrol 53:2497–2522. doi: 10.1093/petrology/egs057
Downes PJ, Dunkley DJ, Fletcher IR, McNaughton NJ, Rasmussen B, Jaques AL, Verrall M,
Sweetapple MT (2016) Zirconolite, zircon and monazite-(Ce) U–Th–Pb age constraints
on the emplacement, deformation and alteration history of the Cummins Range
carbonatite complex, Halls Creek Orogen, Kimberley region, Western Australia. Miner
Petrol 110:199–222. doi: 10.1007/s00710-015-0418-y
Drake H, Tullborg E-L, MacKenzie AB (2009) Detecting the near-surface redox front in
crystalline bedrock using fracture mineral distribution, geochemistry and U-series
disequilibrium. Appl Geochem 24:1023–1039. doi: 10.1016/j.apgeochem.2009.03.004
Dziggel A, Poujol M, Otto A, Kisters AFM, Trieloff M, Schwarz WH, Meyer FM (2010)
New U–Pb and 40Ar/39Ar ages from the northern margin of the Barberton greenstone belt,
South Africa: implications for the formation of Mesoarchaean gold deposits. Precambrian
Res 179:206–220. doi: 10.1016/j.precamres.2010.03.006
Eichhorn R, Schärer U, Höll R (1995) Age and evolution of scheelite-hosting rocks in the
Felbertal deposit (Eastern Alps): U–Pb geochronology of zircon and titanite. Contrib
Mineral Petrol 119:377–386. doi: 10.1007/BF00286936
England GL, Rasmussen B, McNaughton NJ, Fletcher IR, Groves DI, Krapež B (2001)
SHRIMP U–Pb ages of diagenetic and hydrothermal xenotime from the Archaean
Witwatersrand Supergroup of South Africa. Terra Nova 13:360–367. doi: 10.1046/j.1365-
49
3121.2001.00363.x
Fayek M, Harrison TM, Grove M, Coath CD (2000) A rapid in situ method for determining
the ages of uranium oxide minerals: evolution of the Cigar Lake deposit, Athabasca
Basin. Int Geol Rev 42:163–171.
Fisher CM, Longerich HP, Jackson SE, Hanchar JM (2010) Data acquisition and calculation
of U–Pb isotopic analyses using laser ablation (single collector) inductively coupled
plasma mass spectrometry. J Anal At Spectrom 25:1905–1920. doi: 10.1039/C004955G
Fletcher IR, McNaughton NJ, Aleinikoff JA, Rasmussen B, Kamo SL (2004) Improved
calibration procedures and new standards for U–Pb and Th–Pb dating of Phanerozoic
xenotime by ion microprobe. Chem Geol 209:295–314. doi:
10.1016/j.chemgeo.2004.06.015
Fletcher IR, McNaughton NJ, Davis WJ, Rasmussen B (2010) Matrix effects and calibration
limitations in ion probe U–Pb and Th–Pb dating of monazite. Chem Geol 270:31–44.
Frei R (1996) The extent of inter-mineral isotope equilibrium: a systematic bulk U–Pb and Pb
step leaching (PbSL) isotope study of individual minerals from the Tertiary granite of
Jerissos (northern Greece). Eur J Mineral 8:1175–1190. doi: 10.1127/ejm/8/5/1175
Frimmel HE, Groves DI, Kirk J, Ruiz J, Chesley J, Minter WEL (2005) The formation and
preservation of the Witwatersrand goldfields, the world's largest gold province. Econ
Geol 100th Anniversary Volume:769–797.
Frimmel HE, Hennigh Q (2015) First whiffs of atmospheric oxygen triggered onset of crustal
gold cycle. Miner Deposita 50:5–23. doi: 10.1007/s00126-014-0574-8
Froude DO, Ireland TR, Kinny PD, Williams IS, Compston W, Williams IR, Myers JS (1983)
Ion microprobe identification of 4,100–4,200 Myr-old terrestrial zircons. Nature
304:616–618. doi: 10.1038/304616a0
Fryer BJ, Jackson SE, Longerich HP (1993) The application of laser ablation microprobe-
inductively coupled plasma-mass spectrometry (LAM-ICP-MS) to in situ (U)–Pb
geochronology. Chem Geol 109:1–8. doi: 10.1016/0009-2541(93)90058-Q
Fu Y, Sun X, Zhou H, Lin H, Yang T (2016) In-situ LA–ICP–MS U–Pb geochronology and
trace elements analysis of polygenetic titanite from the giant Beiya gold–polymetallic
deposit in Yunnan Province, Southwest China. Ore Geol Rev 77:43–56.
Gardiner NJ, Hawkesworth CJ, Robb LJ, Whitehouse MJ, Roberts NMW, Kirkland CL,
Evans NJ (2017) Contrasting granite metallogeny through the zircon record: a case study
from Myanmar. Sci Rep 7:748. doi: 10.1038/s41598-017-00832-2
Gehrels GE, Valencia VA, Ruiz J (2008) Enhanced precision, accuracy, efficiency, and
spatial resolution of U–Pb ages by laser ablation–multicollector–inductively coupled
plasma–mass spectrometry. Geochem Geophys Geosyst 9:Q03017. doi:
10.1029/2007GC001805
50
Gelcich S, Davis DW, Spooner ETC (2005) Testing the apatite-magnetite geochronometer:
U–Pb and 40Ar/39Ar geochronology of plutonic rocks, massive magnetite-apatite tabular
bodies, and IOCG mineralization in northern Chile. Geochim Cosmochim Acta 69:3367–
3384. doi: 10.1016/j.gca.2004.12.020
Grandia F, Asmerom Y, Getty S, Cardellach E, Canals A (2000) U–Pb dating of MVT ore-
stage calcite: implications for fluid flow in a Mesozoic extensional basin from Iberian
Peninsula. J Geochem Explor 69-70:377–380. doi: 10.1016/S0375-6742(00)00030-3
Gregory CJ, Rubatto D, Allen CM, Williams IS, Hermann J, Ireland T (2007) Allanite micro-
geochronology: a LA-ICP-MS and SHRIMP U–Th–Pb study. Chem Geol 245:162–182.
doi: 10.1016/j.chemgeo.2007.07.029
Griffin WL, Batumike JM, Greau Y, Pearson NJ, Shee SR, O'Reilly SY (2014) Emplacement
ages and sources of kimberlites and related rocks in southern Africa: U–Pb ages and Sr–
Nd isotopes of groundmass perovskite. Contrib Mineral Petrol 168:1032. doi:
10.1007/s00410-014-1032-4
Guillong M, Horn I, Günther D (2003) A comparison of 266 nm, 213 nm and 193 nm
produced from a single solid state Nd:YAG laser for laser ablation ICP-MS. J Anal At
Spectrom 18:1224–1230. doi: 10.1039/B305434A
Gulson BL, Jones MT (1992) Cassiterite: potential for direct dating of mineral deposits and a
precise age for the Bushveld Complex granites. Geology 20:355–358. doi: 10.1130/0091-
7613(1992)020<0355:CPFDDO>2.3.CO;2
Harley SL, Kelly NM (2007) Zircon tiny but timely. Elements 3:13–18. doi:
10.2113/gselements.3.1.13
Harris AC, Allen CM, Bryan SE, Campbell IH, Holcombe RJ, Palin MJ (2004) ELA-ICP-MS
U–Pb zircon geochronology of regional volcanism hosting the Bajo de la Alumbrera Cu-
Au deposit: implications for porphyry-related mineralization. Miner Deposita 39:46–67.
doi: 10.1007/s00126-003-0381-0
51
Harris AC, Dunlap WJ, Reiners PW, Allen CM, Cooke DR, White NC, Campbell IH,
Golding SD (2008) Multimillion year thermal history of a porphyry copper deposit:
application of U–Pb, 40Ar/39Ar and (U–Th)/He chronometers, Bajo de la Alumbrera
copper-gold deposit, Argentina. Miner Deposita 43:295–314. doi: 10.1007/s00126-007-
0151-5
Heaman L (2003) The timing of kimberlite magmatism in North America: implications for
global kimberlite genesis and diamond exploration. Lithos 71:153–184.
Heaman LM (2009) The application of U–Pb geochronology to mafic, ultramafic and alkaline
rocks: an evaluation of three mineral standards. Chem Geol 261:43–52. doi:
10.1016/j.chemgeo.2008.10.021
Heaman LM, Pell J, Grütter HS, Creaser RA (2015) U–Pb geochronology and Sr/Nd isotope
compositions of groundmass perovskite from the newly discovered Jurassic Chidliak
kimberlite field, Baffin Island, Canada. Earth Planet Sci Lett 415:183–199. doi:
10.1016/j.epsl.2014.12.056
Hedenquist JW, Lowenstern JB (1994) The role of magmas in the formation of hydrothermal
ore deposits. Nature 370:519–527. doi: 10.1038/370519a0
Heinrich CA (2015) Witwatersrand gold deposits formed by volcanic rain, anoxic rivers and
Archaean life. Nature Geosci 8:206–209. doi: 10.1038/ngeo2344
Hinthorne JR, Andersen CA, Conrad RL, Lovering JF (1979) Single-grain 207Pb/206Pb and
U/Pb age determinations with a 10-µm spatial resolution using the ion microprobe mass
analyzer (IMMA). Chem Geol 25:271–303. doi: 10.1016/0009-2541(79)90061-5
Hinton RW, Long JVP (1979) High-resolution ion-microprobe measurement of lead isotopes:
variations within single zircons from Lac Seul, northwestern Ontario. Earth Planet Sci
Lett 45:309–325. doi: 10.1016/0012-821X(79)90132-8
Hofmann BA, Knill MD (1996) Geochemistry and genesis of the Lengenbach Pb-Zn-As-Tl-
Ba-mineralisation, Binn Valley, Switzerland. Miner Deposita 31:319–339. doi:
10.1007/BF02280795
Holden NE (1990) Total half-lives for selected nuclides. Pure Appl Chem 62:941–958. doi:
10.1351/pac199062050941
Holmes A (1913) The age of the Earth. Harper and Brothers, London and New York
52
Holmes A (1911) The association of lead with uranium in rock-minerals, and its application
to the measurement of geological time. Proc. R. Soc. Lond. A 85:248–256. doi:
10.1098/rspa.1911.0036
Holmes A (1954) The oldest dated minerals of the Rhodesian Shield. Nature 173:612–614.
Horn I, Rudnick RL, McDonough WF (2000) Precise elemental and isotope ratio
determination by simultaneous solution nebulization and laser ablation-ICP-MS:
application to U–Pb geochronology. Chem Geol 164:281–301.
Horstwood MSA, Košler J, Gehrels G, Jackson SE, McLean NM, Paton C, Pearson NJ,
Sircombe K, Sylvester P, Vermeesch P, Bowring JF, Condon DJ, Schoene B (2016)
Community-derived standards for LA-ICP-MS U–(Th–)Pb geochronology - uncertainty
propagation, age interpretation and data reporting. Geostand Geoanal Res 40:311–332.
doi: 10.1111/j.1751-908X.2016.00379.x
Hoskin PWO, Schaltegger U (2003) The composition of zircon and igneous and metamorphic
petrogenesis. Rev Mineral Geochem 53:27–62. doi: 10.2113/0530027
Huston DL, Maas R, Cross A, Hussey KJ, Mernagh TP, Fraser G, Champion DC (2016) The
Nolans Bore rare-earth element-phosphorus-uranium mineral system: geology, origin and
post-depositional modifications. Miner Deposita 51:797–822. doi: 10.1007/s00126-015-
0631-y
Ireland TR (2014) Ion microscopes and microprobes. In: Treatise on Geochemistry. Elsevier,
pp 385–409
Jaffey AH, Flynn KF, Glendenin LE, Bentley WC, Essling AM (1971) Precision
measurement of half-lives and specific activities of 235U and 238U. Phys Rev C 4:1889–
1906.
Jung S, Mezger K (2003) U–Pb garnet chronometry in high-grade rocks – case studies from
the central Damara orogen (Namibia) and implications for the interpretation of Sm–Nd
garnet ages and the role of high U–Th inclusions. Contrib Mineral Petrol 146:382–396.
doi: 10.1007/s00410-003-0506-6
Kelly NM, Harley SL (2005) An integrated microtextural and chemical approach to zircon
geochronology: refining the Archaean history of the Napier Complex, east Antarctica.
Contrib Mineral Petrol 149:57–84. doi: 10.1007/s00410-004-0635-6
Kerrich R, Kyser TK (1994) 100 Ma timing paradox of Archean gold, Abitibi greenstone belt
(Canada): new evidence from U–Pb and Pb–Pb evaporation ages of hydrothermal zircons.
Geology 22:1131–1134. doi: 10.1130/0091-7613(1994)022<1131:MTPOAG>2.3.CO;2
53
Klemme S, Meyer H-P (2003) Trace element partitioning between baddeleyite and
carbonatite melt at high pressures and high temperatures. Chem Geol 199:233–242.
Klötzli U, Klötzli E, Günes Z, Košler J (2009) Accuracy of laser ablation U–Pb zircon dating:
results from a test using five different reference zircons. Geostand Geoanal Res 33:5–15.
doi: 10.1111/j.1751-908X.2009.00921.x
Koglin N, Zeh A, Frimmel HE, Gerdes A (2010) New constraints on the auriferous
Witwatersrand sediment provenance from combined detrital zircon U–Pb and Lu–Hf
isotope data for the Eldorado Reef (Central Rand Group, South Africa). Precambrian Res
183:817–824.
Kositcin N, Krapež B (2004) Relationship between detrital zircon age-spectra and the tectonic
evolution of the Late Archaean Witwatersrand Basin, South Africa. Precambrian Res
129:141–168. doi: 10.1016/j.precamres.2003.10.011
Kositcin N, McNaughton NJ, Griffin BJ, Fletcher IR, Groves DI, Rasmussen B (2003)
Textural and geochemical discrimination between xenotime of different origin in the
Archaean Witwatersrand Basin, South Africa. Geochim Cosmochim Acta 67:709–731.
doi: 10.1016/S0016-7037(02)01169-9
Košler J, Sláma J, Belousova E, Corfú F, Gehrels GE, Gerdes A, Horstwood MSA, Sircombe
KN, Sylvester PJ, Tiepolo M, Whitehouse MJ, Woodhead JD (2013) U–Pb detrital zircon
analysis – results of an inter-laboratory comparison. Geostand and Geoanal Res 37:243–
259. doi: 10.1111/j.1751-908X.2013.00245.x
Krogh TE (1982) Improved accuracy of U–Pb zircon ages by the creation of more concordant
systems using an air abrasion technique. Geochim Cosmochim Acta 46:637–649. doi:
10.1016/0016-7037(82)90165-X
54
Krogh TE, Davis GL (1975) The production and preparation of 205Pb for use as a tracer for
isotope dilution analyses. Carnegie Institute of Washington, Yearbook 74:416–417.
Kröner A, Kovach VP, Kozakov IK, Kirnozova T, Azimov P, Wong J, Geng HY (2015)
Zircon ages and Nd–Hf isotopes in UHT granulites of the Ider Complex: a cratonic
terrane within the Central Asian Orogenic Belt in NW Mongolia. Gondwana Res
27:1392–1406. doi: 10.1016/j.gr.2014.03.005
Lancelot J, Vitrac A, Allegre CJ (1976) Uranium and lead isotopic dating with grain-by-grain
zircon analysis: a study of complex geological history with a single rock. Earth Planet Sci
Lett 29:357–366. doi: 10.1016/0012-821X(76)90140-0
Larsen ES, Keevil NB, Harrison HC (1952) Method for determining the age of igneous rocks
using the accessory minerals. Geol Soc Am Bull 63:1045–1052. doi: 10.1130/0016-
7606(1952)63[1045:MFDTAO]2.0.CO;2
Law JDM, Phillips GN (2005) Hydrothermal replacement model for Witwatersrand gold.
Econ Geol 100th Anniversary Volume:799–811.
Li C-Y, Zhang R-Q, Ding X, Ling M-X, Fan W-M, Sun W-D (2016) Dating cassiterite using
laser ablation ICP-MS. Ore Geol Rev 72:313–322. doi: 10.1016/j.oregeorev.2015.07.016
Li J-W, Deng X-D, Zhou M-F, Liu Y-S, Zhao X-F, Guo J-L (2010) Laser ablation ICP-MS
titanite U–Th–Pb dating of hydrothermal ore deposits: A case study of the Tonglushan
Cu–Fe–Au skarn deposit, SE Hubei Province, China. Chem Geol 270:56–67. doi:
10.1016/j.chemgeo.2009.11.005
Li XH, Su L, Chung SL, Li ZX, Liu Y, Song B, Liu DY (2005) Formation of the Jinchuan
ultramafic intrusion and the world's third largest Ni-Cu sulfide deposit: associated with
the ∼825 Ma south China mantle plume? Geochem Geophys Geosyst 6:Q11004. doi:
10.1029/2005GC001006
Lobato LM, Santos JOS, McNaughton NJ, Fletcher IR, Noce CM (2007) U–Pb SHRIMP
monazite ages of the giant Morro Velho and Cuiabá gold deposits, Rio das Velhas
greenstone belt, Quadrilátero Ferrífero, Minas Gerais, Brazil. Ore Geol Rev 32:674–680.
55
doi: 10.1016/j.oregeorev.2006.11.007
Ludwig KA, Shen C-C, Kelley DS, Cheng H, Edwards RL (2011) U–Th systematics and
230
Th ages of carbonate chimneys at the Lost City hydrothermal field. Geochim
Cosmochim Acta 75:1869–1888. doi: 10.1016/j.gca.2011.01.008
Ludwig KR (1980) Calculation of uncertainties of U–Pb isotope data. Earth Planet Sci Lett
46:212–220.
Ludwig KR (2012) Isoplot 3.75, a geochronological toolkit for Microsoft Excel. Berkeley
Geochronology Center Special Publication 75.
Luo J-C, Hu R-Z, Fayek M, Li C-S, Bi X-W, Abdu Y, Chen Y-W (2015) In-situ SIMS
uraninite U–Pb dating and genesis of the Xianshi granite-hosted uranium deposit, South
China. Ore Geol Rev 65:968–978. doi: 10.1016/j.oregeorev.2014.06.016
Luvizotto GL, Zack T, Meyer HP, Ludwig T, Triebold S, Kronz A, Münker C, Stöckli DF,
Prowatke S, Klemme S, Jacob DE, von Eynatten H (2009) Rutile crystals as potential
trace element and isotope mineral standards for microanalysis. Chem Geol 261:346–369.
doi: 10.1016/j.chemgeo.2008.04.012
Mattinson JM (2013) Revolution and evolution: 100 years of U–Pb geochronology. Elements
9:53–57. doi: 10.2113/gselements.9.1.53
McKinney ST, Cottle JM, Lederer GW (2015) Evaluating rare earth element (REE)
mineralization mechanisms in Proterozoic gneiss, Music Valley, California. Geol Soc Am
Bull 127:1135–1152. doi: 10.1130/B31165.1
McLean NM, Bowring JF, Bowring SA (2011) An algorithm for U–Pb isotope dilution data
reduction and uncertainty propagation. Geochem Geophys Geosyst 12:Q0AA18. doi:
10.1029/2010GC003478
McLean NM, Bowring JF, Gehrels G (2016) Algorithms and software for U–Pb
geochronology by LA-ICPMS. Geochem Geophys Geosyst 17. doi:
56
10.1002/2015GC006097
McLean NM, Condon DJ, Schoene B, Bowring SA (2015) Evaluating uncertainties in the
calibration of isotopic reference materials and multi-element isotopic tracers
(EARTHTIME Tracer Calibration Part II). Geochim Cosmochim Acta 164:481–501. doi:
10.1016/j.gca.2015.02.040
Meier DL, Heinrich CA, Watts MA (2009) Mafic dikes displacing Witwatersrand gold reefs:
evidence against metamorphic-hydrothermal ore formation. Geology 37:607–610. doi:
10.1130/G25657A.1
Meinhold G (2010) Rutile and its applications in earth sciences. Earth Sci Rev 102:1–28. doi:
10.1016/j.earscirev.2010.06.001
Melleton J, Gloaguen E, Frei D, Novák M, Breiter K (2012) How are the emplacement of
rare-element pegmatites, regional metamorphism and magmatism interrelated in the
Moldanubian domain of the Variscan Bohemian Massif, Czech Republic? Can Min
50:1751–1773. doi: 10.3749/canmin.50.6.1751
Mezger K, Hanson GN, Bohlen SR (1989) U–Pb systematics of garnet: dating the growth of
garnet in the late Archean Pikwitonei granulite domain at Cauchon and Natawahunan
lakes, Manitoba, Canada. Contrib Mineral Petrol 101:136–148. doi: 10.1007/BF00375301
Michael Meyer F, Kolb J, Sakellaris GA, Gerdes A (2006) New ages from the Mauritanides
Belt: recognition of Archean IOCG mineralization at Guelb Moghrein, Mauritania. Terra
Nova 18:345–352. doi: 10.1111/j.1365-3121.2006.00698.x
Michard-Vitrac A, Lancelot J, Allegre CJ, Moorbath S (1977) U–Pb ages on single zircons
from the early Precambrian rocks of West Greenland and the Minnesota River Valley.
Earth Planet Sci Lett 35:449–453. doi: 10.1016/0012-821X(77)90077-2
Miller JS, Matzel JEP, Miller CF, Burgess SD, Miller RB (2007) Zircon growth and recycling
during the assembly of large, composite arc plutons. J Volcanol Geotherm Res 167:282–
299. doi: 10.1016/j.jvolgeores.2007.04.019
Moreto CPN, Monteiro LVS, Xavier RP, Creaser RA, DuFrane SA, Melo GHC, da Silva
MAD, Tassinari CCG, Sato K (2014) Timing of multiple hydrothermal events in the iron
oxide–copper–gold deposits of the Southern Copper Belt, Carajás Province, Brazil. Miner
Deposita 50:517–546. doi: 10.1007/s00126-014-0549-9
Morisset C-E, Scoates JS, Weis D, Friedman RM (2009) U–Pb and 40Ar/39Ar geochronology
of the Saint-Urbain and Lac Allard (Havre-Saint-Pierre) anorthosites and their associated
Fe–Ti oxide ores, Québec: evidence for emplacement and slow cooling during the
57
collisional Ottawan Orogeny in the Grenville Province. Precambrian Res 174:95–116.
doi: 10.1016/j.precamres.2009.06.009
Moser DE (1997) Dating the shock wave and thermal imprint of the giant Vredefort impact,
South Africa. Geology 25:7–10. doi: 10.1130/0091-
7613(1997)025<0007:DTSWAT>2.3.CO;2
Moser DE, Davis WJ, Reddy SM, Flemming RL, Hart RJ (2009) Zircon U–Pb strain
chronometry reveals deep impact-triggered flow. Earth Planet Sci Lett 277:73–79. doi:
10.1016/j.epsl.2008.09.036
Mosoh Bambi CK, Frimmel HE, Zeh A, Suh CE (2013) Age and origin of Pan-African
granites and associated U–Mo mineralization at Ekomédion, southwestern Cameroon. J
Afr Earth Sci 88:15–37. doi: 10.1016/j.jafrearsci.2013.08.005
Möller A, O’Brien PJ, Kennedy A, Kröner A (2003) Linking growth episodes of zircon and
metamorphic textures to zircon chemistry: an example from the ultrahigh-temperature
granulites of Rogaland (SW Norway). Geol Soc Spec Pub 220:65–81. doi:
10.1144/GSL.SP.2003.220.01.04
Mueller AG, Campbell IH, Schiotte L, Sevigny JH, Layer PW (1996) Constraints on the age
of granitoid emplacement, metamorphism, gold mineralization, and subsequent cooling of
the Archean greenstone terrane at Big Bell, Western Australia. Econ Geol 91:896–915.
doi: 10.2113/gsecongeo.91.5.896
Mueller AG, Hall GC, Nemchin AA, Stein HJ, Creaser RA, Mason DR (2007) Archean high-
Mg monzodiorite–syenite, epidote skarn, and biotite–sericite gold lodes in the Granny
Smith–Wallaby district, Australia: U–Pb and Re–Os chronometry of two intrusion-related
hydrothermal systems. Miner Deposita 43:337–362. doi: 10.1007/s00126-007-0164-0
Muhling JR, Fletcher IR, Rasmussen B (2012) Dating fluid flow and Mississippi Valley type
base-metal mineralization in the Paleoproterozoic Earaheedy Basin, Western Australia.
Precambrian Res 212-213:75–90. doi: 10.1016/j.precamres.2012.04.016
Muntean JL, Frimmel HE, Phillips N, Law J, Myers R (2005) Controversies on the origin of
world-class gold deposits, Part II: Witwatersrand gold deposits. Soc Econ Geol
Newsletter 60:7–12–19.
Müller SG, Krapež B, Barley ME, Fletcher IR (2005) Giant iron-ore deposits of the
Hamersley province related to the breakup of Paleoproterozoic Australia: new insights
from in situ SHRIMP dating of baddeleyite from mafic intrusions. Geology 33:577–580.
doi: 10.1130/G21482.1
Norcross C, Davis DW, Spooner E, Rust A (2000) U–Pb and Pb–Pb age constraints on
Paleoproterozoic magmatism, deformation and gold mineralization in the Omai area,
Guyana Shield. Precambrian Res 102:69–86. doi: 10.1016/s0301-9268(99)00102-3
Oberli F, Meier M, Berger A, Rosenberg CL, Gieré, R (2004) U–Th–Pb and 230Th/238U
58
disequilibrium isotope systematics: precise accessory mineral chronology and melt
evolution tracing in the Alpine Bergell intrusion. Geochim Cosmochim Acta 68:2543–
2560. doi: 10.1016/j.gca.2003.10.017
Ono S, Fayek M (2011) Decoupling of O and Pb isotope systems of uraninite in the early
Proterozoic conglomerates in the Elliot Lake district. Chem Geol 288:1–13. doi:
10.1016/j.chemgeo.2010.03.015
Pal DC, Chaudhuri T, McFarlane C, Mukherjee A, Sarangi AK (2011) Mineral chemistry and
in situ dating of allanite, and geochemistry of its host rocks in the Bagjata uranium mine,
Singhbhum Shear Zone, India—implications for the chemical evolution of REE
mineralization and mobilization. Econ Geol 106:1155–1171. doi:
10.2113/econgeo.106.7.1155
Paquette JL, Tiepolo M (2007) High resolution (5 µm) U–Th–Pb isotope dating of monazite
with excimer laser ablation (ELA)-ICPMS. Chem Geol 240:222–237. doi:
10.1016/j.chemgeo.2007.02.014
Parrish RR (1990) U–Pb dating of monazite and its application to geological problems. Can J
Earth Sci 27:1431–1450. doi: 10.1139/e90-152
Parrish RR, Noble SR (2003) Zircon U–Th–Pb geochronology by isotope dilution – thermal
ionization mass spectrometry (ID-TIMS). Rev Mineral Geochem 53:183–213. doi:
10.2113/0530183
Peterman EM, Mattinson JM, Hacker BR (2012) Multi-step TIMS and CA-TIMS monazite
U–Pb geochronology. Chem Geol 312-313:58–73.
Petersson J, Whitehouse MJ, Eliasson T (2001) Ion microprobe U–Pb dating of hydrothermal
xenotime from an episyenite: evidence for rift-related reactivation. Chem Geol 175:703–
712. doi: 10.1016/S0009-2541(00)00338-7
Pettke T, Oberli F, Heinrich CA (2010) The magma and metal source of giant porphyry-type
ore deposits, based on lead isotope microanalysis of individual fluid inclusions. Earth
Planet Sci Lett 296:267–277. doi: 10.1016/j.epsl.2010.05.007
59
Pfaff K, Romer RL, Markl G (2009) U–Pb ages of ferberite, chalcedony, agate, “U-mica” and
pitchblende: constraints on the mineralization history of the Schwarzwald ore district. Eur
J Mineral 21:817–836. doi: 10.1127/0935-1221/2009/0021-1944
Philippe S, Lancelot JR, Clauer N, Pacquet A (2011) Formation and evolution of the Cigar
Lake uranium deposit based on U–Pb and K–Ar isotope systematics. Can J Earth Sci
30:720–730. doi: 10.1139/e93-058
Phillips GN, Law JDM (2000) Witwatersrand gold fields; geology, genesis, and exploration.
Rev Econ Geol 13:439–500.
Pigois J-P, Groves DI, Fletcher IR, McNaughton NJ, Snee LW (2003) Age constraints on
Tarkwaian palaeoplacer and lode-gold formation in the Tarkwa-Damang district, SW
Ghana. Miner Deposita 38:695–714. doi: 10.1007/s00126-003-0360-5
Poletti JE, Cottle JM, Hagen-Peter GA, Lackey JS (2016) Petrochronological constraints on
the origin of the Mountain Pass ultrapotassic and carbonatite intrusive suite, California. J
Petrol 57:1555–1598. doi: 10.1093/petrology/egw050
Polito PA, Kyser TK, Thomas D, Marlatt J, Drever G (2005) Re-evaluation of the
petrogenesis of the Proterozoic Jabiluka unconformity-related uranium deposit, Northern
Territory, Australia. Miner Deposita 40:257–288. doi: 10.1007/s00126-005-0007-9
Poujol M (2007) An overview of the Pre-Mesoarchean rocks of the Kaapvaal Craton, South
Africa. Developments in Precambrian Geology 15:453–463. doi: 10.1016/S0166-
2635(07)15051-9
Prowatke S, Klemme S (2006) Trace element partitioning between apatite and silicate melts.
Geochim Cosmochim Acta 70:4513–4527. doi: 10.1016/j.gca.2006.06.162
Rao NVC, Wu F-Y, Mitchell RH, Li Q-L, Lehmann B (2013) Mesoproterozoic U–Pb ages,
trace element and Sr–Nd isotopic composition of perovskite from kimberlites of the
eastern Dharwar craton, southern India: distinct mantle sources and a widespread 1.1 Ga
tectonomagmatic event. Chem Geol 353:48–64. doi: 10.1016/j.chemgeo.2012.04.023
Rasbury ET, Cole JM (2009) Directly dating geologic events: U–Pb dating of carbonates. Rev
Geophys 47:RG3001. doi: 10.1029/2007RG000246
Rasmussen B, Blake TS, Fletcher IR (2005) U–Pb zircon age constraints on the Hamersley
spherule beds: evidence for a single 2.63 Ga Jeerinah-Carawine impact ejecta layer.
Geology 33:725–728. doi: 10.1130/G21616.1
Rasmussen B, Fletcher IR, Muhling JR, Mueller AG, Hall GC (2007a) Bushveld-aged fluid
flow, peak metamorphism, and gold mobilization in the Witwatersrand basin, South
Africa: constraints from in situ SHRIMP U–Pb dating of monazite and xenotime.
60
Geology 35:931–934. doi: 10.1130/G23588A.1
Rasmussen B, Fletcher IR, Muhling JR, Thorne WS, Broadbent GC (2007b) Prolonged
history of episodic fluid flow in giant hematite ore bodies: evidence from in situ U–Pb
geochronology of hydrothermal xenotime. Earth Planet Sci Lett 258:249–259. doi:
10.1016/j.epsl.2007.03.033
Rasmussen B, Mueller AG, Fletcher IR (2008) Zirconolite and xenotime U–Pb age
constraints on the emplacement of the Golden Mile Dolerite sill and gold mineralization
at the Mt Charlotte mine, Eastern Goldfields Province, Yilgarn Craton, Western
Australia. Contrib Mineral Petrol 157:559–572. doi: 10.1007/s00410-008-0352-7
Richards JP (2013) Giant ore deposits formed by optimal alignments and combinations of
geological processes. Nature Geosci 6:911–916. doi: 10.1038/ngeo1920
Robert J, Miranda CF, Muxart R (1969) Mesure de la période du protactinium 231 par
microcalorimétrie. Radiochimica Acta 11:104–108. doi: 10.1524/ract.1969.11.2.104
Romer RL (1992) Vesuvianite – new tool for the U–Pb dating of skarn ore deposits. Miner
Petrol 46:331–341.
Romer RL, Lüders V (2006) Direct dating of hydrothermal W mineralization: U–Pb age for
hübnerite (MnWO4), Sweet Home Mine, Colorado. Geochim Cosmochim Acta 70:4725–
4733. doi: 10.1016/j.gca.2006.07.003
Romer RL, Martinsson O, Perdahl JA (1994) Geochronology of the Kiruna iron ores and
hydrothermal alterations. Econ Geol 89:1249–1261. doi: 10.2113/gsecongeo.89.6.1249
61
Romer RL, Öhlander B (1994) U–Pb age of the Yxsjöberg tungsten-skarn deposit, Sweden.
GFF 116:161–166. doi: 10.1080/11035899409546179
Romer RL, Smeds S-A (1996) U–Pb columbite ages of pegmatites from Sveconorwegian
terranes in southwestern Sweden. Precambrian Res 76:15–30. doi: 10.1016/0301-
9268(95)00023-2
Romer RL, Smeds S-A (1994) Implications of U–Pb ages of columbite-tantalites from
granitic pegmatites for the Palaeoproterozoic accretion of 1.90–1.85 Ga magmatic arcs to
the Baltic Shield. Precambrian Res 67:141–158. doi: 10.1016/0301-9268(94)90008-6
Romer RL, Smeds SA, Černý P (1996) Crystal-chemical and genetic controls of U–Pb
systematics of columbite-tantalite. Miner Petrol 57:243–260. doi: 10.1007/BF01162361
Romer RL, Wright JE (1992) U–Pb dating of columbites: a geochronologic tool to date
magmatism and ore deposits. Geochim Cosmochim Acta 56:2137–2142. doi:
10.1016/0016-7037(92)90337-I
Ruiz J, Valencia VA, Chesley JT, Kirk J, Gehrels G, Frimmel HE (2006) The source of gold
for the Witwatersrand from Re–Os and U–Pb detrital zircon geochronology. Geochim
Cosmochim Acta 70:A543.
Rutherford E (1929) Origin of actinium and age of the Earth. Nature 123:313–314.
Salier BP, Groves DI, McNaughton NJ, Fletcher IR (2005) Geochronological and stable
isotope evidence for widespread orogenic gold mineralization from a deep-seated fluid
source at ca 2.65 Ga in the Laverton gold province, Western Australia. Econ Geol
100:1363–1388. doi: 10.2113/gsecongeo.100.7.1363
Salier BP, Groves DI, McNaughton NJ, Fletcher IR (2004) The world-class Wallaby gold
deposit, Laverton, Western Australia: an orogenic-style overprint on a magmatic-
hydrothermal magnetite-calcite alteration pipe? Miner Deposita 39:473–494. doi:
10.1007/s00126-004-0425-0
Sarma DS, Fletcher IR, Rasmussen B, McNaughton NJ, Mohan MR, Groves DI (2011)
Archaean gold mineralization synchronous with late cratonization of the Western
Dharwar Craton, India: 2.52 Ga U–Pb ages of hydrothermal monazite and xenotime in
gold deposits. Miner Deposita 46:273–288. doi: 10.1007/s00126-010-0326-3
62
Schaltegger U, Pettke T, Audétat A, Reusser E, Heinrich CA (2005) Magmatic-to-
hydrothermal crystallization in the W–Sn mineralized Mole Granite (NSW, Australia).
Chem Geol 220:215–235. doi: 10.1016/j.chemgeo.2005.02.018
Schaltegger U, Schmitt AK, Horstwood MSA (2015) U–Th–Pb zircon geochronology by ID-
TIMS, SIMS, and laser ablation ICP-MS: recipes, interpretations, and opportunities.
Chem Geol 402:89–110. doi: 10.1016/j.chemgeo.2015.02.028
Schärer U (1984) The effect of initial 230Th disequilibrium on young U–Pb ages: the Makalu
case, Himalaya. Earth Planet Sci Lett 67:191–204. doi: 10.1016/0012-821x(84)90114-6
Schärer U, Corfu F, Demaiffe D (1997) U–Pb and Lu–Hf isotopes in baddeleyite and zircon
megacrysts from the Mbuji-Mayi kimberlite: constraints on the subcontinental mantle.
Chem Geol 143:1–16. doi: 10.1016/S0009-2541(97)00094-6
Schmitt AK, Zack T (2012) High-sensitivity U–Pb rutile dating by secondary ion mass
spectrometry (SIMS) with an O2+ primary beam. Chem Geol 332-333:65–73. doi:
10.1016/j.chemgeo.2012.09.023
Schmitz MD, Bowring SA (2001) U–Pb zircon and titanite systematics of the Fish Canyon
Tuff: an assessment of high-precision U–Pb geochronology and its application to young
volcanic rocks. Geochim Cosmochim Acta 65:2571–2587.
Schoene B, Crowley JL, Condon DJ, Schmitz MD, Bowring SA (2006) Reassessing the
uranium decay constants for geochronology using ID-TIMS U–Pb data. Geochim
Cosmochim Acta 70:426–445. doi: 10.1016/j.gca.2005.09.007
Seo J, Choi S-G, Kim DW, Park J-W, Oh CW (2015) A new genetic model for the Triassic
Yangyang iron-oxide–apatite deposit, South Korea: constraints from in situ U–Pb and
trace element analyses of accessory minerals. Ore Geol Rev 70:110–135. doi:
63
10.1016/j.oregeorev.2015.04.009
Simon AC, Ripley EM (2011) The role of magmatic sulfur in the formation of ore deposits.
Rev Mineral Geochem 73:513–578. doi: 10.2138/rmg.2011.73.16
Singer DA, Berger VI, Moring BC (2008) Porphyry copper deposits of the world: Database
and grade and tonnage models, 2008. US Geological Survey
Skirrow RG, Bastrakov EN, Barovich K, Fraser GL, Creaser RA, Fanning CM, Raymond OL,
Davidson GJ (2007) Timing of iron oxide Cu-Au-(U) hydrothermal activity and Nd
isotope constraints on metal sources in the Gawler Craton, South Australia. Econ Geol
102:1441–1470.
Skirrow RG, Mercadier J, Armstrong R, Kuske T, Deloule É (2016) The Ranger uranium
deposit, northern Australia: timing constraints, regional and ore-related alteration, and
genetic implications for unconformity-related mineralisation. Ore Geol Rev 76:463–503.
doi: 10.1016/j.oregeorev.2015.09.001
Skolodowska Curie M (1898) Rayons émis par les composés de l’uranium et du thorium. C R
Acad Sci 126:1101–1103.
Smith CB, Allsopp HL, Garvie OG, Kramers JD, Jackson PFS, Clement CR (1989) Note on
the U–Pb perovskite method for dating kimberlites: examples from the Wesselton and De
Beers mines, South Africa, and Somerset Island, Canada. Chem Geol 79:137–145. doi:
10.1016/0168-9622(89)90016-X
Smith MP, Storey CD, Jeffries TE, Ryan C (2009) In situ U–Pb and trace element analysis of
accessory minerals in the Kiruna district, Norrbotten, Sweden: new constraints on the
timing and origin of mineralization. J Petrol 50:2063–2094. doi:
10.1093/petrology/egp069
Smith SR, Foster GL, Romer RL, Tindle AG, Kelley SP, Noble SR, Horstwood M, Breaks
FW (2004) U–Pb columbite-tantalite chronology of rare-element pegmatites using TIMS
and laser ablation-multi collector-ICP-MS. Contrib Mineral Petrol 147:549–564. doi:
10.1007/s00410-003-0538-y
64
Smye AJ, Roberts NMW, Condon DJ, Horstwood MSA, Parrish RR (2014) Characterising
the U–Th–Pb systematics of allanite by ID and LA-ICPMS: implications for
geochronology. Geochim Cosmochim Acta 135:1–28. doi: 10.1016/j.gca.2014.03.021
Spencer CJ, Kirkland CL, Taylor RJM (2016) Strategies towards statistically robust
interpretations of in situ U–Pb zircon geochronology. Geosci Front 7:581–589. doi:
10.1016/j.gsf.2015.11.006
Stacey JS, Kramers JD (1975) Approximation of terrestrial lead isotope evolution by a two-
stage model. Earth Planet Sci Lett 26:207–221. doi: 10.1016/0012-821X(75)90088-6
Stelten ME, Cooper KM, Vazquez JA, Calvert AT, Glessner JJG (2015) Mechanisms and
timescales of generating eruptible rhyolitic magmas at Yellowstone Caldera from zircon
and sanidine geochronology and geochemistry. J Petrol 56:1607–1642. doi:
10.1093/petrology/egv047
Storey CD, Jeffries TE, Smith M (2006) Common lead-corrected laser ablation ICP–MS U–
Pb systematics and geochronology of titanite. Chem Geol 227:37–52.
Storey CD, Smith MP, Jeffries TE (2007) In situ LA-ICP-MS U–Pb dating of metavolcanics
of Norrbotten, Sweden: records of extended geological histories in complex titanite
grains. Chem Geol 240:163–181. doi: 10.1016/j.chemgeo.2007.02.004
Stosch H-G, Romer RL, Daliran F, Rhede D (2010) Uranium–lead ages of apatite from iron
oxide ores of the Bafq District, East-Central Iran. Miner Deposita 46:9–21. doi:
10.1007/s00126-010-0309-4
Tallarico FHB, McNaughton NJ, Groves DI, Fletcher IR, Figueiredo BR, Carvalho JB, Rego
JL, Nunes AR (2004) Geological and SHRIMP II U–Pb constraints on the age and origin
of the Breves Cu-Au-(W-Bi-Sn) deposit, Carajás, Brazil. Miner Deposita 39:68–86. doi:
10.1007/s00126-003-0383-y
Tapster S, Condon DJ, Naden J, Noble SR, Petterson MG, Roberts NMW, Saunders AD,
Smith DJ (2016) Rapid thermal rejuvenation of high-crystallinity magma linked to
porphyry copper deposit formation; evidence from the Koloula Porphyry Prospect,
Solomon Islands. Earth Planet Sci Lett 442:206–217. doi: 10.1016/j.epsl.2016.02.046
Taylor RD, Goldfarb RJ, Monecke T, Fletcher IR, Cosca MA, Kelly NM (2015) Application
of U–Th–Pb phosphate geochronology to young orogenic gold deposits: new age
constraints on the formation of the Grass Valley gold district, Sierra Nevada Foothills
Province, California. Econ Geol 110:1313–1337. doi: 10.2113/econgeo.110.5.1313
Tera F, Wasserburg GJ (1972a) U–Th–Pb systematics in lunar highland samples from the
Luna 20 and Apollo 16 missions. Earth Planet Sci Lett 17:36–51.
65
Tera F, Wasserburg GJ (1972b) U–Th–Pb systematics in three Apollo 14 basalts and the
problem of initial Pb in lunar rocks. Earth Planet Sci Lett 14:281–304. doi: 10.1016/0012-
821X(72)90128-8
Tilton GR (1960) Volume diffusion as a mechanism for discordant lead ages. J Geophys Res:
Solid Earth 65:2933–2945. doi: 10.1029/JZ065i009p02933
Tilton GR, Davis GL, Wetherill GW, Aldrich LT (1957) Isotopic ages of zircon from granites
and pegmatites. Eos T Am Geophys Un 38:360–371. doi: 10.1029/TR038i003p00360
Tilton GR, Patterson C, Brown H, Inghram M, Hayden R, Hess D, Larsen E (1955) Isotopic
composition and distribution of lead, uranium, and thorium in a Precambrian granite.
Geol Soc Am Bull 66:1131–1148. doi: 10.1130/0016-
7606(1955)66[1131:ICADOL]2.0.CO;2
Torrealday HI, Hitzman MW, Stein HJ, Markley RJ, Armstrong R, Broughton D (2000) Re–
Os and U–Pb dating of the vein-hosted mineralization at the Kansanshi copper deposit,
northern Zambia. Econ Geol 95:1165–1170. doi: 10.2113/gsecongeo.95.5.1165
Valley PM, Fisher CM, Hanchar JM, Lam R, Tubrett M (2010) Hafnium isotopes in zircon: a
tracer of fluid-rock interaction during magnetite–apatite (“Kiruna-type”) mineralization.
Chem Geol 275:208–220. doi: 10.1016/j.chemgeo.2010.05.011
Vallini DA, Groves DI, McNaughton NJ, Fletcher IR (2006) Uraniferous diagenetic xenotime
in northern Australia and its relationship to unconformity-associated uranium
mineralisation. Miner Deposita 42:51–64. doi: 10.1007/s00126-005-0012-z
Vermeesch P (2012) On the visualisation of detrital age distributions. Chem Geol 312-
313:190–194. doi: 10.1016/j.chemgeo.2012.04.021
Vielreicher NM, Groves DI, Snee LW, Fletcher IR, McNaughton NJ (2010) Broad
synchroneity of three gold mineralization styles in the Kalgoorlie Gold Field: SHRIMP,
66
U–Pb, and 40Ar/39Ar geochronological evidence. Econ Geol 105:187–227. doi:
10.2113/gsecongeo.105.1.187
von Grosse A (1932) On the origin of the actinium series of radioactive elements. Phys Rev
42:565–570. doi: 10.1103/PhysRev.42.565
von Quadt A, Gallhofer D, Guillong M, Peytcheva I, Waelle M, Sakata S (2014) U–Pb dating
of CA/non-CA treated zircons obtained by LA-ICP-MS and CA-TIMS techniques:
impact for their geological interpretation. J Anal At Spectrom 29:1618–1629. doi:
10.1039/C4JA00102H
Vry JK, Baker JA (2006) LA-MC-ICPMS Pb–Pb dating of rutile from slowly cooled
granulites: confirmation of the high closure temperature for Pb diffusion in rutile.
Geochim Cosmochim Acta 70:1807–1820. doi: 10.1016/j.gca.2005.12.006
Wall CJ, Scoates JS (2016) High-precision U–Pb zircon-baddeleyite dating of the J-M reef
platinum group element deposit in the Stillwater Complex, Montana (USA). Econ Geol
111:771–782. doi: 10.2113/econgeo.111.3.771
Wan B, Xiao W, Zhang L, Han C (2012) Iron mineralization associated with a major strike–
slip shear zone: Radiometric and oxygen isotope evidence from the Mengku deposit, NW
China. Ore Geol Rev 44:136–147. doi: 10.1016/j.oregeorev.2011.09.011
Watson EB, Chemiak DJ, Hanchar JM, Harrison TM, Wark DA (1997) The incorporation of
Pb into zircon. Chem Geol 141:19–31. doi: 10.1016/S0009-2541(97)00054-5
Webber GR, Hurley PM, Fairbairn HW (1956) Relative ages of eastern Massachusetts
granites by total lead ratios in zircon. Am J Sci 254:574–583. doi: 10.2475/ajs.254.9.574
Webster JD, Piccoli PM (2015) Magmatic apatite: a powerful, yet deceptive, mineral.
Elements 11:177–182. doi: 10.2113/gselements.11.3.177
Weis P, Driesner T, Heinrich CA (2012) Porphyry-copper ore shells form at stable pressure-
temperature fronts within dynamic fluid plumes. Science 338:1613–1616. doi:
10.1126/science.1225009
Wendt I (1984) A three-dimensional U–Pb discordia plane to evaluate samples with common
lead of unknown isotopic composition. Chem Geol 46:1–12. doi: 10.1016/0009-
67
2541(84)90162-1
Wendt I, Carl C (1991) The statistical distribution of the mean squared weighted deviation.
Chem Geol 86:275–285. doi: 10.1016/0168-9622(91)90010-T
White LT, Ireland TR (2012) High-uranium matrix effect in zircon and its implications for
SHRIMP U–Pb age determinations. Chem Geol 306-307:78–91. doi:
10.1029/2011GC003726
Wingate MTD, Compston W (2000) Crystal orientation effects during ion microprobe U–Pb
analysis of baddeleyite. Chem Geol 168:75–97. doi: 10.1016/S0009-2541(00)00184-4
Woodhead JD, Hergt JM, Simonson BM (1998) Isotopic dating of an Archean bolide impact
horizon, Hamersley basin, Western Australia. Geology 26:47–50. doi: 10.1130/0091-
7613(1998)026<0047:IDOAAB>2.3.CO;2
Wotzlaw J-F, Bindeman IN, Watts KE, Schmitt AK, Caricchi L, Schaltegger U (2014)
Linking rapid magma reservoir assembly and eruption trigger mechanisms at evolved
Yellowstone-type supervolcanoes. Geology 42:807–810.
Wotzlaw J-F, Bindeman IN, Stern RA, D’Abzac F-X, Schaltegger U (2015) Rapid
heterogeneous assembly of multiple magma reservoirs prior to Yellowstone
supereruptions. Sci Rep 5:14026. doi: 10.1038/srep14026
Wotzlaw JF, Schaltegger U, Frick DA, Dungan MA, Gerdes A, Günther D (2013) Tracking
the evolution of large-volume silicic magma reservoirs from assembly to supereruption.
Geology 41:867–870. doi: 10.1130/G34366.1
Wu F-Y, Arzamastsev AA, Mitchell RH, Li Q-L, Sun J, Yang Y-H, Wang R-C (2013a)
Emplacement age and Sr–Nd isotopic compositions of the Afrikanda alkaline ultramafic
complex, Kola Peninsula, Russia. Chem Geol 353:210–229. doi:
10.1016/j.chemgeo.2012.09.027
Wu F-Y, Mitchell RH, Li Q-L, Sun J, Liu C-Z, Yang Y-H (2013b) In situ U–Pb age
determination and Sr–Nd isotopic analysis of perovskite from the Premier (Cullinan)
kimberlite, South Africa. Chem Geol 353:83–95. doi: 10.1016/j.chemgeo.2012.06.002
Wu F-Y, Yang Y-H, Li Q-L, Mitchell RH, Dawson JB, Brandl G, Yuhara M (2011) In situ
determination of U–Pb ages and Sr–Nd–Hf isotopic constraints on the petrogenesis of the
Phalaborwa carbonatite Complex, South Africa. Lithos 127:309–322.
Yang W-B, Niu H-C, Shan Q, Sun W-D, Zhang H, Li N-B, Jiang Y-H, Yu X-Y (2013)
Geochemistry of magmatic and hydrothermal zircon from the highly evolved Baerzhe
alkaline granite: implications for Zr–REE–Nb mineralization. Miner Deposita 49:451–
470. doi: 10.1007/s00126-013-0504-1
York D (1968) Least squares fitting of a straight line with correlated errors. Earth Planet Sci
Lett 5:320–324. doi: 10.1016/S0012-821X(68)80059-7
68
Yuan S, Peng J, Hao S, Li H, Geng J, Zhang D (2011) In situ LA-MC-ICP-MS and ID-TIMS
U–Pb geochronology of cassiterite in the giant Furong tin deposit, Hunan Province, South
China: New constraints on the timing of tin–polymetallic mineralization. Ore Geol Rev
43:235–242. doi: 10.1016/j.oregeorev.2011.08.002
Zack T, Stockli DF, Luvizotto GL, Barth MG, Belousova E, Wolfe MR, Hinton RW (2011)
In situ U–Pb rutile dating by LA-ICP-MS: 208Pb correction and prospects for geological
applications. Contrib Mineral Petrol 162:515–530. doi: 10.1007/s00410-011-0609-4
Zartman RE, Smith JV (2009) Mineralogy and U–Th–Pb age of a uranium-bearing jasperoid
vein, Sunshine Mine, Coeur d'Alene district, Idaho, USA. Chem Geol 261:185–195.
doi:10.1016/j.chemgeo.2008.09.006
Zeh A, Ovtcharova M, Wilson AH, Schaltegger U (2015) The Bushveld Complex was
emplaced and cooled in less than one million years – results of zirconology, and
geotectonic implications. Earth Planet Sci Lett 418:103–114. doi:
10.1016/j.epsl.2015.02.035
Zhang D, Peng J, Coulson IM, Hou L, Li S (2014) Cassiterite U–Pb and muscovite 40Ar–39Ar
age constraints on the timing of mineralization in the Xuebaoding Sn–W–Be deposit,
western China. Ore Geol Rev 62:315–322. doi: 10.1016/j.oregeorev.2014.04.011
Zi J-W, Rasmussen B, Muhling JR, Fletcher IR, Thorne AM, Johnson SP, Cutten HN,
Dunkley DJ, Korhonen FJ (2015) In situ U–Pb geochronology of xenotime and monazite
from the Abra polymetallic deposit in the Capricorn Orogen, Australia: dating
hydrothermal mineralization and fluid flow in a long-lived crustal structure. Precambrian
Res 260:91–112. doi: 10.1016/j.precamres.2015.01.010
69