You are on page 1of 13

Radiometric Dating by Rb-Sr, Sm-Nd, Lu-Hf, Re-Os, and Pb-Pb

Jeff Vervoort, School of the Environment, Washington State University, Pullman, WA, United States
© 2020 Elsevier Inc. All rights reserved.

Introduction 1
Fundamentals 1
Radiogenic Isotope Systems 3
Rb-Sr 3
Sm-Nd 4
Lu-Hf 7
Re-Os 8
Pb-Pb 11
References 12

Glossary
BSE Bulk Silicate Earth. The solid Earth not including the core.
CHUR Chondritic Uniform Reservoir. Best determined from unprocessed, unequilibrated, chondritic meteorites. Generally
assumed to be the best estimate of Bulk Silicate Earth (BSE) for refractory, lithophile elements.
Ga Giga annum (billion years).
Granulite A metamorphic rock formed at high temperature and pressure.
Lithophile Referring to the affinity of an element for the silicate part of the Earth (crust and mantle).
Ma Mega annum (million years).
Orogen A general term to describe a belt of Earth’s crust involved in compressional tectonic activity (i.e., mountain belt
formation).
TIMS Thermal Ionization Mass Spectrometry.
VMS Volcanogenic Massive Sulfide. An ore deposit containing Cu, Ni, Pb that forms in concert with the eruption of
submarine volcanic rocks.
Z Atomic number. This is also the number of protons that define an element.

Introduction

All of the systems for radiometric dating discussed in this chapter have long half-lives, with most measured in the billions of years.
As a result, they are used to determine the age of events on Earth on geological time frames of tens of millions to billions of
years. In sum, these methods have helped to provide an early understanding of the ages of rocks that make up our continents and
the geological events that have modified them through time. More recently, their role in determining the absolute ages of rocks and
events on Earth has largely been supplanted by other geochronologic methods such as U-Pb zircon and 40Ar-39Ar geochronology as
these provide generally more precise and accurate age constraints. Nevertheless, the methods in this chapter still play an important
role in geochronology in more specialized applications as discussed below.

Fundamentals
All geochronology, including the systems discussed in this chapter, is based on the fundamental equation of radioactive decay:
dN
¼ − lN, (1)
dt
which states that the change in the number of radioactive atoms (dN) with time (dt) is proportional to the number of atoms
originally present (N) and the decay constant (l). The negative sign on l indicates that radioactive atoms are decreasing in number
with time. From this expression, the radiogenic decay equation can be derived

N ¼ N0  e −lt , (2)
which states that the number of radioactive parent isotopes (N) present at a given time, t, is a function of the number present
initially (N0), the decay constant for that radionuclide, and time.
The radioactive decay of a parent isotope also produces a radiogenic daughter isotope (D∗), which is equal to the number of
parent decays: D∗ ¼ N0 − N. From this we get the expression for the production of radiogenic daughters from a radioactive parent:

Encyclopedia of Geology, 2nd edition https://doi.org/10.1016/B978-0-08-102908-4.00038-2 1


2 Radiometric Dating by Rb-Sr, Sm-Nd, Lu-Hf, Re-Os, and Pb-Pb

 
D ∗ ¼ N  elt − 1 : (3)

For most rocks and minerals in the isotope systems in this chapter, there are pre-existing daughter isotopes (D0) that need to be
taken into account:
 
D ¼ D0 + N  elt − 1 : (4)

In practice, the parent and daughter isotopes are measured relative to a stable (i.e., not radioactive or radiogenic) isotope of the
daughter element, Dr
  
D D P  
¼ +  elt − 1 : (5)
Dr p Dr i Dr p

As written here, i denotes the initial isotope composition (e.g., at the time of crystallization or formation age), p signifies the present
time, D/Dr is the isotope ratio of the daughter element, and P/Dr is the parent/daughter ratio (ratio of the radioactive parent isotope
to the stable daughter reference isotope). This equation, however, has two unknowns: time, t, and the initial ratio, R(i), making it
impossible to determine an age from just this expression because the initial isotope ratio, R(i), cannot be known independently.
In these cases, the isochron method can be used to determine an age. This approach uses the isotope composition of a suite of rocks
or minerals that are assumed to have had the same isotope composition at the time of their simultaneous formation. This can then
be expressed in two parallel equations for phases A and B:
  
D D P  
¼ +  elt − 1 , (6)
Dr AðpÞ Dr AðiÞ Dr AðpÞ

and
  
D D P  
¼ +  elt − 1 : (7)
Dr BðpÞ Dr BðiÞ Dr Bð p Þ

The isochron approach assumes that all phases are in isotopic equilibrium at the time of their formation. If this is true, D/Dr
A(i ) ¼ D/Dr B(i ) and therefore, when taking the difference between these two equations, the initial isotope composition of both
phases drops out. This leaves the expression:
  "   #
D D P P  
− ¼ −  elt − 1 : (8)
Dr AðpÞ Dr BðpÞ Dr AðpÞ Dr BðpÞ

Defining the difference in isotopic composition between the two phases as D, and rearranging terms gives

D
D
Dr ðpÞ  
 ¼ elt − 1 : (9)
P
D
Dr ðpÞ

The expression on the left hand of the equation defines a slope on a D/Dr vs. P/Dr diagram as shown in Fig. 1.

Fig. 1 Diagram illustrating the principle of the isochron method.


Radiometric Dating by Rb-Sr, Sm-Nd, Lu-Hf, Re-Os, and Pb-Pb 3

Starting with the simple case of a crystallizing homogeneous magma, the different mineral phases forming in that rock at t ¼ 0
will have the same daughter isotope (D/Dr) composition (A in Fig. 1) but, because of the different partitioning of elements into
these phases, will have different P/Dr ratios (B). Over time, the daughter isotope accumulates at the expense of the parent due to its
radioactive decay, thus resulting in an increase in D/Dr ratios with time (C). A phase with more parent relative to the reference
isotope (i.e., higher P/Dr) will evolve to higher D/Dr. At some later time, t,—if the system has been remained closed to elemental
disturbance—these different mineral phases will define a line (D) whose slope is proportional to the time since the system was last
homogenized. In practice, multiple points are used to define the line and slope, which help to identify inherited components that
were not in isotopic equilibrium with the rest of the rock, and also to identify open-system behavior.
If a slope can be determined for a set of phases with the same starting isotope composition (and assuming a closed system), an
isochron age can be calculated using the expression:
 
slope ¼ elt − 1 , (10)

which can be rearranged to give:

t ¼ ln ðslope + 1Þ=l: (11)


Thus, from this method it is possible not only to determine an age, but also an initial isotope composition from the intercept of the
regression line, i.e., the point where the parent/daughter ratio ¼ 0 and, therefore, the isotope composition is time invariant (A in
Fig. 1).
There are several requirements necessary for the isochron method to yield a reliable age: (1) All components in the system (i.e.,
those that are included as points on the isochron) must have formed at the same time, t0; (2) all system components must have the
same isotope composition when they formed (i.e., the system was isotopically homogenous at t0); and (3) the isotope system was
not subsequently disturbed during later geological events since t0 (there was neither loss nor gain of either parent or daughter
element for any point on the isochron; i.e., a closed system). In reality, not all of these requirements may be entirely satisfied in the
strictest sense, but deviations must be minimal to achieve an accurate and meaningful age. In addition, P/Dr ratios and D/Dr isotope
variations need to be significant relative to their ability to be measured in order to determine a well-defined slope and, by extension,
a precise age.

Radiogenic Isotope Systems


Rb-Sr
The Rb-Sr geochronometer is based on the decay of 87Rb (27.8% abundance) to 87Sr (7.0% abundance) by a b− decay, with a half
life of 49.6 billion years (4.9624  1010 year). The corresponding decay constant is 1.3968  10−11 year−1.
87
Rb ! 87 Sr +b − : (12)
Rb (atomic number, Z ¼ 37) is an alkali element with a valence of +1. It is very soluble in water and, as a result, Rb is one of the most
mobile elements. Rb, owing to its valence and large ionic radius, is also one of the most highly incompatible elements (concentrated
in the liquid relative to the solid phase during melting) and, therefore, is concentrated in Earth’s crust over the mantle. Sr (Z ¼ 38) is
an alkali-earth element with a valence of +2. Like Rb, it is soluble resulting in it also being a mobile element, but less so than Rb. Sr is
also an incompatible element and similarly concentrated in the crust, but less so than Rb. As a result, the crust has both higher Rb
and Sr concentrations and higher Rb/Sr ratios than the mantle.
Rb substitutes freely with K, due to their geochemical similarities as Group 1 alkali elements and, therefore, is particularly
concentrated in K-bearing minerals such as micas and potassium feldspars. Similarly, Sr substitutes commonly with Ca as both are
Group 2 alkali-earth elements and is concentrated in Ca-rich minerals such as plagioclase feldspar, calcite, and apatite.
By convention, the reference isotope for the Rb-Sr isotope system is the stable isotope 86Sr; the resulting radiogenic equation is,
87  87  87 
 
Sr Sr Rb
86 ¼ +  elt − 1 : (13)
Sr p 86 Sr i 86 Sr p

This equation states that the 87Sr/86Sr isotope ratio today (p) is a function of the ratio of 87Sr/86Sr initially (i) plus the amount added
by radiogenic production of 87Sr, which, in turn, is a function of the Rb/Sr ratio, the decay constant of 87Rb, and the amount of
elapsed time (since time i). The higher the Rb/Sr ratio and the longer the time, the higher (more radiogenic) the 87Sr/86Sr ratio
becomes. These systematics, in turn, can be used to calculate an age using the isochron method, where D87Sr/86Sr/D87Rb/86Sr is the
slope on an Rb-Sr isochron diagram (e.g., Fig. 2).
20 87 Sr  1 3
D 86 Sr
t ¼ ln 4 @ 87
 + 15=l87Rb :
A (14)
Rb
D 86 Sr

There are clear strengths and weaknesses of the Rb-Sr method for geochronology. One advantage is that both Rb and Sr are fairly
common minor elements and are relatively abundant in most rocks and minerals. Second, a variety of petrogenetic processes can
4 Radiometric Dating by Rb-Sr, Sm-Nd, Lu-Hf, Re-Os, and Pb-Pb

Fig. 2 Rb-Sr isochron for the Amîtsoq gneisses, SW Greenland. After Moorbath S, Allaart JH, Bridgewater D and McGregor VR (1977) Rb-Sr ages of early Archaean
supracrustal rocks and Amitsog gneisses at Isua. Nature 270: 43–45.

fractionate Rb/Sr ratios in minerals leading to a wide range of Rb/Sr ratios that are necessary for determining a robust isochron.
Conversely, because Sr—and especially Rb—are elements that are easily mobilized under a range of geologic conditions, the Rb-Sr
isotope system is often compromised and does not satisfy the requirements for a closed isotopic system. As a result, its current use in
geochronology for providing robust primary ages is limited.
Historically, the Rb-Sr method was used to provide some of the earliest ages that helped us to understand the chronology of the
Earth. This was because Rb and Sr isotope analyses were some of the first that could be done accurately and precisely using thermal
ionization mass spectrometry (TIMS), the cutting-edge technology for measuring isotope ratios at the time. Shown in Fig. 2 is a
whole-rock Rb-Sr isochron diagram for the Amîtsoq gneiss of southwest Greenland determined by Moorbath et al. (1977). At the
time, this was the oldest dated rock and demonstrated the existence of crust of great antiquity on Earth (and SW Greenland, in
particular). This isochron, however, illustrates some of the limitations of the Rb-Sr geochronometer. Although this sample has a
wide range of Rb/Sr ratios, it is composed of whole-rock samples that likely have (i) different ages and (ii) different initial isotopic
compositions, thereby violating two of the fundamental requirements necessary for an isochron to produce an accurate age.
In addition, because of the mobility of Sr and, especially, Rb, it is likely that this isochron has not been a closed system since the
rocks formed. Therefore, this isochron probably represents a rough approximation of a regional metamorphic event or events—the
time(s) when Rb and Sr were widely remobilized—rather than the time of rock formation. These limitations are now widely
recognized and Rb-Sr geochronology, which was widely used in the 1970s and 1980s, has been replaced by other geochronometers
and is now rarely used as a dating technique. This method, however, can still have utility in dating minimally disturbed systems,
such as meteorites, and for determining mineral isochrons in pristine, closed-system terrestrial rocks.
An example of the latter is illustrated by a study of the Great Dyke of Zimbabwe (Nebel and Mezger, 2008). Five biotite separates
with leached ortho-and clinopyroxene separates from two proximal pyroxenites, yield a precise Rb-Sr age of 2543.0  4.4 Ma
(Fig. 3). Note that the slope of the isochron for this sample—and, therefore, the age—is controlled by the very high Rb/Sr of the
biotite separates (87Sr/86Sr from 330 to over 1020). This age is 32 Ma younger than the crystallization age of the Great Dyke,
determined by U-Pb zircon geochronology. The younger Rb-Sr isochron date was interpreted to represent a cooling age due to the
lower closure temperature for biotite (Nebel and Mezger, 2008).

Sm-Nd
The Sm-Nd geochronometer is based on the decay of 147Sm (15.0%) to 143Nd (12.2%) by a-decay, with a long half-life of 106
billion years (1.06  1011 year). The corresponding decay constant is 6.54  10−12 year−1.
147
Sm ! 143 Nd + a (15)
Radiometric Dating by Rb-Sr, Sm-Nd, Lu-Hf, Re-Os, and Pb-Pb 5

Fig. 3 Rb-Sr isochron for the Great Dyke, Zimbabwe. Modified after Nebel O and Mezger K (2008) Timing of thermal stabilization of the Zimbabwe craton deduced
from high-precision Rb-Sr chronology, great dyke. Precambrian Research 164: 227–232.

Both Sm (Z ¼ 62) and Nd (Z ¼ 60) are light to middle rare-earth elements (REE) with a valence of +3 and small ionic radii
(Sm ¼ 0.96 Å; Nd ¼ 0.99 Å). The relatively high ratios of valence to ionic radius for these elements makes both moderately
incompatible and therefore, concentrated in the crust over the mantle. The REEs, as a group, have very similar geochemical behavior
with the largest difference among them being the slight, but systematic, decreasing ionic radius with increasing atomic number.
The consequence of this is that the Sm/Nd ratio is not greatly fractionated by most magmatic processes and Sm/Nd variations
between common rocks and minerals tend to be relatively small.
Sm and Nd both have low solubilities and are considered immobile under most conditions. The REEs are trace elements in most
crustal materials and have abundance levels in the ppm (Sm) to 10s of ppm (Nd) range. Some REEs can be accommodated by—and
therefore concentrated in—certain accessory and metamorphic minerals such as apatite, monazite, sphene, and garnet. Garnet, for
example, accommodates REEs with an increasing affinity with increasing atomic number (i.e., garnet accommodates Sm more than
Nd). This generally results in much higher (and variable) 147Sm/144Nd ratios in garnet relative to the bulk rock and this feature
can be utilized in Sm-Nd geochronology.
By convention, the reference isotope for the Sm-Nd system is 144Nd (23.8% abundance); the resulting radiogenic equation is:
! ! !
143
Nd 143
Nd 147
Sm  
144 ¼ +  elt − 1 (16)
Nd p 144 Nd i 144 Nd p

An age can be determined using this system from the slope on an isochron diagram (e.g., Fig. 4) using the equation:
20 143 Nd  1 3
D 144 Nd
t ¼ ln 4@ 147  A + 15=l147Sm (17)
Sm
D 144 Nd

The strengths of the Sm-Nd method include the relative immobility of the parent-daughter elements, which help to maintain a
closed system for these rocks, and the ability to measure the isotope compositions of these elements precisely and accurately.
In addition, certain accessory and metamorphic phases (e.g., garnet) effectively fractionate Sm/Nd, which can result in higher and
variable, parent/daughter ratios necessary to produce a well-defined isochron. In contrast, the very similar geochemical behavior of
Sm and Nd in most rock-forming minerals results in small variations in Sm/Nd which are insufficient for precise geochronology.
As is true for Rb-Sr, the Sm-Nd method was used in the early years of the development of this system, particularly in the late 1970s
and 1980s. The current use of the Sm-Nd geochronometer is mostly restricted to specialized applications like garnet geochronology
and, to a lesser extent, for mineral isochrons.
An example of one of the early Sm-Nd isochrons is shown in Fig. 4 and illustrates some of the limitations of this method as a
robust chronometer for bulk-rock compositions. The samples plotted in this diagram are volcanic rocks from the Archean
Onverwacht Group that have a large range in composition. Overall, the samples have 147Sm/144Nd ratios that vary by a factor of
two, but that is achieved only by including felsic samples with very low 147Sm/144Nd. Excluding those samples, the total variation in
147
Sm/144Nd is 25%, insufficient for a meaningful isochron.
This example illustrates the dilemma of Sm-Nd dating for bulk rock compositions and major rock-forming minerals: magmatic
processes do not significantly fractionate Sm/Nd ratios, resulting in a very limited spread in Sm/Nd and, thus, poorly-defined,
imprecise, isochrons. In order to produce a larger spread and a more precise isochron, other compositions are often included in the
6 Radiometric Dating by Rb-Sr, Sm-Nd, Lu-Hf, Re-Os, and Pb-Pb

Fig. 4 Sm-Nd isochron for a package of volcanic rocks from the Onverwacht Group, South Africa. After Hamilton PJ, Evensen, NM, O’Nions RK, Smith HS and
Erlank AJ (1979) Sm-Nd dating of Onverwacht group volcanics, southern Africa. Nature 279: 298–300.

isochron, as was the case in Fig. 4. Doing this, however, likely violates two of the fundamental requirements for the isochron
method: (i) the components of the isochron must have all formed at the same time; and (ii) they must have had the same isotope
composition at the time they were formed. It is unlikely these requirements can be met for a mixed package of rocks.
Sm-Nd geochronology still has utility in meteoritic samples where there is sufficient Sm/Nd fractionation and in mineral
isochrons with a high Sm/Nd phase, most notably in garnet geochronology (Fig. 5). In particular, garnet geochronology has seen
increasing use in recent years because it holds the promise of directly dating a common index mineral produced during
metamorphic reactions. Fig. 5 shows an isochron from a garnet bearing paragenesis. Garnet has highly elevated Sm/Nd ratios
and controls the slope of the isochron. The bulk rock has low Sm/Nd ratios, typically less than bulk-Earth estimates (e.g., less than
the 147Sm/144Nd of CHUR, chondritic uniform reservoir, 0.1960; Bouvier et al., 2008). These characteristics have been used to
produce relatively precise and robust Sm-Nd ages and are being used increasingly in concert with Lu-Hf garnet, and monazite U-Pb,
geochronology to more fully understand metamorphic histories.

Fig. 5 Garnet Sm-Nd isochron. Note that the four garnet fractions have highly elevated Sm/Nd ratios relative to the whole rock. Johnson TA, Vervoort JD, Ramsey
MJ, Southworth S and Mulcahy SR (2020) Tectonic evolution of the Grenville Orogen in the central Appalachians. Precambrian Research 346: 105740.
Radiometric Dating by Rb-Sr, Sm-Nd, Lu-Hf, Re-Os, and Pb-Pb 7

Lu-Hf
The Lu-Hf geochronometer is based on the b− decay of 176Lu (2.6% abundance) to 176Hf (5.3% abundance), with a half-life of 37.1
billion years (3.71  1010 year). The corresponding decay constant is 1.867  10−11 year−1.
176
Lu ! 176 Hf + b − (18)
In terms of geochemical characteristics, Lu is the heaviest of the REEs (Z ¼ 71) with a valence of +3 and with the smallest ionic
radius of the REEs. As a result, it is moderately incompatible, but the least incompatible of the REEs. In other respects, it behaves
similarly to the other REEs, especially adjacent Yb (Z ¼ 70), which is why this group of elements often occur together in nature and
why they are not easily separated in the lab. Lu, like the other REEs, is insoluble and considered to be among the most immobile of
elements. Lu occurs with the other REEs, mostly in accessory phases. As Lu is the heaviest of the REEs, it is especially accommodated
in those phases that prefer the HREEs. Garnet is an excellent example of this and most compositions strongly partition Lu relative to
the lighter REEs and Hf.
The adjacent heavier element to Lu, Hf (Z ¼ 72), is not a REE, but is characterized as a high field strength element (HFSE), because
of its high charge (+4 valence) to size (0.71 Å) ratio. Hf has very close geochemical similarities to Zr and, to a lesser extent Ti, as these
are fellow group IVB elements and have identical charge and similar ionic radii. Hf can substitute for both but especially with Zr. As a
result, Hf is highly concentrated in zircon (1% Hf in zircon), baddeleyite, and to a lesser extent, ilmenite, rutile, and sphene.
Despite the different geochemical characteristics of Lu and Hf, these elements are not strongly fractionated from each other in
most crustal rocks—usually no more than a factor of two (and generally significantly less). Most crustal rocks have very low
176
Lu/177Hf ratios, significantly less than CHUR (i.e., 176Lu/177Hf < 0.0336; Bouvier et al., 2008). Variations between highly
evolved and primitive compositions are small, the former with significantly lower Lu/Hf ratios. Some phases, as discussed above
(e.g., garnet, apatite) can concentrate Lu and result in highly elevated Lu/Hf ratios, which then can be exploited for geochronology.
By convention, the reference isotope for the Lu-Hf system is 177Hf (18.6% abundance); the resulting radiogenic equation is:
! ! !
176
Hf 176
Hf 176
Lu  
177 ¼ +  elt − 1 (19)
Hf p 177 Hf i 177 Hf p

An age can be determined from the slope on an isochron diagram (e.g., Fig. 6) using the equation:
20 176  1 3
Hf
D 177 Hf
6B pC 7
t ¼ ln 4@ 176  A + 15=l176Lu (20)
Lu
D 177 Hf
p

Because of the largely parallel behavior of the Lu-Hf and Sm-Nd isotope systems, the strengths and weaknesses are mostly the same.
Both Lu and Hf are largely immobile and can generally maintain closed systems. In addition, certain phases, notably garnet, can
accommodate Lu during growth and result in very high Lu/Hf ratios, which, in turn, can be exploited for use in geochronology.
However, for most rock-forming minerals, and crustal rocks in general, Lu-Hf fractionation is not sufficient for useful
geochronology.

Fig. 6 Eucrite Lu-Hf isochron. After Patchett PJ (1983) Importance of the Lu-Hf isotope system in studies of planetary chronology and chemical evolution.
Geochimica et Cosmochimica Acta 47: 81–91.
8 Radiometric Dating by Rb-Sr, Sm-Nd, Lu-Hf, Re-Os, and Pb-Pb

The first published Lu-Hf isochron by Patchett, 1983 (Fig. 6) was used not to determine an age, but to determine the 176Lu decay
constant, using a variation of Eq. (11):

l ¼ ln ðslope + 1Þ=t: (21)


This isochron used data from eucrite meteorites to determine a slope and an assumed age of 4.55 Ga for eucrite formation. From
this, Patchett determined a value for l176Lu of 1.94  0.07  10−11 year−1. This isochron, however, used data from both basaltic and
mafic cumulate eucrites to achieve sufficient spread to define a slope, something that was not possible with either the basaltic or
cumulate samples by themselves. The l176Lu decay constant was revised to the current value of l176Lu ¼ 1.867  10−11 year−1 in
two age-comparative studies by Scherer et al. (2001) and Söderlund et al. (2004).
The lack of significant Lu/Hf fractionation in crustal rocks and in common rock-forming minerals has prevented the Lu-Hf
isotope system from being used in determining isochron ages for most geologic materials. Large Lu/Hf fractionations, however, are
possible in a few notable phases (e.g., garnet, apatite, lawsonite), and this has been the basis of current geochronology using this
method. The garnet isochron, shown in Fig. 7, illustrates the utility of this method. Because of the high affinity for Lu—and its
exclusion of Hf—garnet can have some extraordinarily high Lu/Hf ratios (150 times chondrite values in Fig. 7), which, in turn,
results in very radiogenic 176Hf/177Hf ratios that can be used to determine precise Lu-Hf isochron dates.
One interesting approach of garnet geochronology is to determine both Sm-Nd and Lu-Hf dates on the same samples, as is
shown in Figs. 5 and 7. In several instances, the dates from the two methods are significantly different and, in some cases, radically
so. This seems to be particularly the case for garnet-bearing granulites from slowly cooled orogens. For example, the Sm-Nd age for
the Virginia Blue Ridge garnet granulite sample C-12-1 (Fig. 5) is over 50 Ma younger than the Lu-Hf age for this same sample. This
clearly represents different behavior in the two isotope systems and is generally interpreted to be due to the difference in closure
temperatures of the Sm-Nd and Lu-Hf systems to elemental diffusion. This comparative approach—and when integrated with other
chronometers (e.g., zircon, monazite, titanite)—can provide a more complete thermal history of metamorphic rocks and terranes.

Re-Os
The Re-Os geochronometer is based on the b− decay of 187Re (62.6% abundance) to 187Os (2.0% abundance), with a half-life of
41.61 Ga (4.161  1010 year). The corresponding decay constant is 1.666  10−11 year−1.
187
Re ! 187 Os + b − (22)
In contrast to the three previous geochronometers described above, all of which involve strongly lithophile elements that are
concentrated in the crust, Re (Z ¼ 75) and Os (Z ¼ 76) are both highly siderophile elements (HSE) concentrated in the core and, as
a result, are depleted in the silicate Earth with very low crustal abundances, typically in the ppb to sub-ppb range. Re is a group VIIB
element, but shows more affinity to Mo rather than the other VIIB element, Mn. Re is particularly concentrated in molybdenite and,
to a lesser extent, in Cu sulfides.
Os, in addition to being a HSE, is also a platinum group element (PGE) and occurs primarily in association with other PGEs. The
abundance of Os in the crust is very low—in the sub-ppb level—but is highest in ultramafic rocks, particularly those with Cu and Ni
sulfides. In mantle phases, Re is moderately incompatible and Os is highly compatible. This results in crustal rocks having highly
elevated Re/Os ratios relative to the mantle. This strong fractionation of the Re/Os ratio—and the existence of phases and some
rocks enriched in Re—forms the basis of Re-Os geochronology.

Fig. 7 Garnet Lu-Hf isochron. Note that the four garnet fractions have highly elevated Lu/Hf ratios relative to the whole rock. Johnson TA, Vervoort JD, Ramsey MJ,
Southworth S and Mulcahy SR (2020) Tectonic evolution of the Grenville Orogen in the central Appalachians. Precambrian Research 346: 105740.
Radiometric Dating by Rb-Sr, Sm-Nd, Lu-Hf, Re-Os, and Pb-Pb 9

Fig. 8 Re-Os isochron for IVB iron meteorites. After Day JMD, Brandon AD and Walker RA (2016) Highly Siderophile elements in earth, Mars, the moon, and
asteroids, reviews in Mineralogy & Geochemistry. Mineralogical Society of America, 161–238.; data from Shen JJ, Papanastassiou DA and Wasserburg GJ (1996)
Precise Re–Os determinations and systematics of iron meteorites. Geochimica et Cosmochimica Acta 60: 2887–2900; and Walker RJ, McDonough WF, Honesto J,
Chabot NL, McCoy TJRDA and Bellucci JJ (2008) Modeling fractional crystallization of group IVB iron meteorites. Geochimica et Cosmochimica Acta 72: 2198–2216.

By convention, the reference isotope for the Re-Os system is 188Os (13.2% abundance); the resulting radiogenic equation is:
  187 
187
Os 187
Os Re  
188 ¼ +  elt − 1 (23)
Os p 188 Os i 188 Os p

If a slope can be determined on an isochron diagram (e.g., Fig. 8), an age can be calculated using the equation:
20 187  1 3
Os
D 188 Os
6B pC 7
t ¼ ln 4@ 187  A + 15=l187Re (24)
Re
D 188 Os
p

Although it is possible to determine an isochron age in this way (e.g., Figs. 8 and 9), because of the very low abundances of Re and
Os in most crustal materials—and the challenges inherent in the chemistry and analysis—these Re-Os isochrons are not very
common. Rather, different strategies have been adopted to take advantage of particular Re-Os behavior and Re abundances as
discussed in the examples below.
One clear strength of the Re-Os method is the very large Re/Os fractionation that occurs during mantle melting and differen-
tiation of those magmas. The very low crustal abundance of Re and Os, however, limits the use of this method mostly to mantle
materials, more primitive crustal compositions, and crustal materials that concentrate Re (molybdenite deposits, black shales).
As is true for the other isochron dating methods in this chapter, the Re-Os method, in practice, is limited to specialized
applications. Although Re/Os fractionations are large during the extraction of melt from the mantle and during crustal melting
and differentiation, the very low concentrations of Re and Os in crustal materials—and the technical challenges of their chemical
separation and analysis—have prevented this method from common use as a chronometer. The applications that have been
successful include: iron meteorite samples enriched in Re and Os; mantle xenoliths with moderate relative abundance of Re and Os;
and crustal samples that are highly enriched in Re and with very high Re/Os ratios.
An example of a Re-Os iron meteorite isochron is shown in Fig. 8. In these samples Os is relatively abundant (ppm to tens of
ppm) allowing for accurate analysis. The restricted spread in Re/Os, however ( a factor of 1.5 in Fig. 8), is not sufficient to generate
a reasonably precise isochron. These iron meteorite samples have not been disturbed by magmatic processes since the closure of the
Re-Os system, at about 4.5 Ga ago.
A different approach with the Re-Os geochronometer is for samples with very high Re/Os ratios such as shown in Fig. 9. In this
example, black shales are highly enriched in Re relative to Os and, while still very low in abundance (Re in the ppb range; Os in the
tens of ppt range) have very high Re/Os ratios (100–1000) and result in a very large dynamic range in 187Os/188Os—sufficient for
producing a precise isochron.
Molybdenite has also been dated using the Re-Os system. Due to its geochemical similarity with Mo, Re is highly concentrated in
molybdenite which displays Re concentrations of tens of ppm. What is different with this approach, however, is that there is
effectively no common Os and therefore all of the Os in the sample is assumed to be radiogenic 187Os. If this assumption is correct
the radiogenic decay equation becomes 187Os ¼ 187Re • (elt–1). An age can be calculated from the expression:
10 Radiometric Dating by Rb-Sr, Sm-Nd, Lu-Hf, Re-Os, and Pb-Pb

Fig. 9 Re-Os isochron for black shales. After Kendall B, Creaser RA and Selby D (2006) Re-Os geochronology of postglacial black shales in Australia: Constraints
on the timing of “Sturtian” glaciation. Geology 34: 729–732.

" ! #
D 187
Os p
t ¼ ln  +1 , (25)
D 187 Re p

using the slope calculated from a plot of 187Re vs. 187Os (Fig. 10), assuming all the 187Os is radiogenic (no common Os) and the
187
Re is in natural abundance.
Finally, mantle xenoliths have also been dated using Re-Os. Most of these xenoliths have sufficient Re and Os to measure, but the
common observation is that the Re/Os ratio has been disturbed by interaction with the host magma during transport of the
xenoliths from the mantle to the Earth’s surface. The approach here is to calculate model ages for these samples by projecting back to
the intersection with a chondritic evolution curve, using either the measured Re/Os of the sample or assuming a Re/Os ratio of zero,
with the assumption that any Re in the sample has been added from the xenolith host magma during transport from the mantle. The
former is called a mantle model age (TMA) and the latter a Re depletion (TRD) age. For a full discussion of this methodology, refer to
Shirey and Walker (1998) and Harvey et al. (2016).

Fig. 10 Re-Os molybdenite dating. After Stein HJ, Markey RJ, Morgan JW, Hannah JL and Schersten A (2001) The remarkable Re-Os chronometer in molybdenite:
How and why it works. Terra Nova 13: 479–486.
Radiometric Dating by Rb-Sr, Sm-Nd, Lu-Hf, Re-Os, and Pb-Pb 11

Pb-Pb
A variant on the isochrons discussed above, has be used in the U-Pb system. The U-Pb system involves two long decay chains, 238U
to 206Pb and 235U to 207Pb with half-lives of 4.47 Ga and 0.707 Ga and decay constants of 1.55125  10−10 year−1 and
9.8571  10−10 year−1, respectively.
238
U ! 206 Pb + 8a + 6b − (26)
207 −
235
U! Pb + 7a + 5b (27)
These can be written in the standard radiogenic equations:
! ! !
206
Pb 206
Pb 238
U  
204 ¼ +  el238U t − 1 (28)
Pb p 204 Pb i 204 Pb p
! ! !
207
Pb 207
Pb 235
U  
204 ¼ 204 + 204  el235U t − 1 (29)
Pb p
Pb i
Pb p

By themselves, these equations are rarely, if ever, used in an isochron application for whole-rock compositions because of the
mobility of U in Earth surface environments. Their utility for isochron dating, however, can be realized by combining and expressing
them only in terms of Pb. Rewriting these equations in terms of the radiogenic Pb added over an interval of time (e.g.,
D207Pb/204Pb ¼ 207Pb/204Pb(p)–207Pb/204Pb(i)), substituting m for the 238U/204Pb ratio, and incorporating the natural 238U/235U
ratio, gives the following expressions:
!
206
Pb  
D 204 ¼ m  el238t − 1 (30)
Pb p
!
207
Pb m  
D 204 ¼  el235t − 1 (31)
Pb 137:818
p

These equations can then be combined to yield the following:


!  l235t 
207
Pb 1 e −1
D 206 ¼  l238t (32)
Pb 137:818 ðe − 1Þ
p

It can be noted that combining Eqs. (30) and (31) eliminates the present-day U/Pb ratios and what remains are only two variables,
D207Pb/206Pb and time. This means that the age (or the 207Pb/206Pb ratio) is not affected by any change in the present-day U/Pb
ratio because it has been removed from the equations. The ratio, D207Pb/206Pb, can be calculated from the slope of a regression line
drawn through a series of cogenetic samples on a 206Pb/204Pb vs. 207Pb/204Pb diagram (Fig. 11) in a way analogous to determining a
Rb-Sr isochron date. Another important difference between this and a conventional isochron is that initial Pb isotopic compositions

Fig. 11 Meteorite Pb-Pb isochron used to determine the first accurate age estimate of the Earth. After Patterson CC (1956) Age of meteorites and the earth.
Geochimica et Cosmochimica Acta 10: 230–237.
12 Radiometric Dating by Rb-Sr, Sm-Nd, Lu-Hf, Re-Os, and Pb-Pb

Fig. 12 Whole rock Pb-Pb isochron. Data from Vervoort JD, White WM, Thorpe RI and Franklin JM (1993) Postmagmatic thermal activity in the Abitibi Greenstone
Belt, Noranda and Matagami districts: Evidence from whole rock Pb isotope data. Economic Geology 88: 1598–1614.

cannot be determined from the intercept method (i.e., when the U/Pb ratio equals zero) because the parent isotopes (235U and
238
U) are no longer in the equation or isochron diagram.
It should also be noted that Eq. (32) is a transcendental equation; time, t, in this equation cannot be solved for directly. Rather,
time is determined iteratively by estimating values for t, calculating a D207Pb/206Pb ratio, and comparing this to the calculated slope
until a solution has been made. 207Pb/206Pb ages are also widely used in zircon geochronology because of the very small amounts of
206
Pb and 207Pb present in zircon when it forms (i.e., “common Pb”), requiring only small corrections. This means that the
207
Pb/206Pb isotope ratios of the zircons can be used to determine ages.
A Pb-Pb isochron for a sequence of volcanic rocks that host rocks volcanic massive sulfide deposits in the Abitibi greenstone belt
in Canada is shown in Fig. 12. The variations in 207Pb/204Pb and 206Pb/204Pb shown here represent the time of the last U/Pb
fractionation in these rocks, in this case, the late Archean at 2642 Ma. This isochron age, however, is 60 Ma younger than the
crystallization age of these rocks and, therefore, probably represents subsequent hydrothermal alteration that affected these rocks
post crystallization.
For samples that have undergone minimal alteration throughout their history, the Pb-Pb method can yield very precise and
accurate ages. An example of this is Pb-Pb geochronology in chondritic meteorites, which has helped to provide a very precise
absolute chronology of early Solar System materials (e.g., Amelin et al., 2010).

References
Amelin Y, Kaltenbach A, Iizuka T, Stirling CH, Ireland T, Petaev M, and Jacobsen SB (2010) U–Pb chronology of the Solar System’s oldest solids with variable 238U/235U. Earth and
Planetary Science Letters 300: 343–350.
Bouvier A, Vervoort JD, and Patchett PJ (2008) The Lu-Hf and Sm-Nd isotopic composition of CHUR: Constraints from unequilibrated chondrites and implications for the bulk
composition of terrestrial planets. Earth and Planetary Science Letters 273: 48–57.
Harvey J, Warren JM, and Shirey SB (2016) Mantle Sulfides and their Role in Re–Os and Pb Isotope Geochronology. Mineralogical Society of America.
Moorbath S, Allaart JH, Bridgewater D, and McGregor VR (1977) Rb-Sr ages of early Archaean supracrustal rocks and Amitsog gneisses at Isua. Nature 270: 43–45.
Nebel O and Mezger K (2008) Timing of thermal stabilization of the Zimbabwe craton deduced from high-precision Rb-Sr chronology, Great Dyke. Precambrian Research
164: 227–232.
Patchett PJ (1983) Importance of the Lu-Hf isotope system in studies of planetary chronology and chemical evolution. Geochimica et Cosmochimica Acta 47: 81–91.
Scherer E, Münker C, and Mezger K (2001) Calibration of the lutetium–hafnium clock. Science 293: 683–686.
Shirey SB and Walker RJ (1998) The Re-Os isotope system in cosmochemistry and high-temperature geochemistry. Annual Review of Earth and Planetary Sciences 26: 423–500.
Söderlund U, Patchett PJ, Vervoort JD, and Isachsen CE (2004) The 176Lu decay constant determined by Lu–Hf and U–Pb isotope systematics of Precambrian mafic intrusions. Earth
and Planetary Science Letters 219: 311–324.

Further reading
Baxter E and Scherer E (2013) Garnet geochronology: Timekeeper of tectonometamorphic processes. Elements 9: 433–438.
Begemann F, Ludwig KR, Lugmair GW, Min K, Nyquist LE, Patchett PJ, Renne P, Shih C-Y, Villa IM, and Walker RJ (2001) Call for an improved set of decay constants for
geochronological use. Geochimia et Cosmochimica Acta 65: 111–121.
Carlson RL (2013) Sm–Nd dating. In: Rink W and Thompson J (eds.) Encyclopedia of Scientific Dating Methods. Dordrecht: Springer.
Day JMD, Brandon AD, and Walker RA (2016) Highly siderophile elements in Earth, Mars, the moon, and asteroids, reviews in mineralogy &amp; geochemistry. Mineralogical Society of
America 161–238.
Radiometric Dating by Rb-Sr, Sm-Nd, Lu-Hf, Re-Os, and Pb-Pb 13

Dickin A (1995) Radiogenic Isotope Geochemistry. Cambridge: Cambridge University Press.


Faure G and Mensing TM (2005) Isotopes: Principles and Applications, 3rd edn. Hoboken: John Wiley & Sons.
Hamilton PJ, Evensen NM, O’Nions RK, Smith HS, and Erlank AJ (1979) Sm-Nd dating of Onverwacht group volcanics, southern Africa. Nature 279: 298–300.
Johnson TA, Vervoort JD, Ramsey MJ, Southworth S, and Mulcahy SR (2020) Tectonic evolution of the Grenville Orogen in the central Appalachians. Precambrian Research 346:
105740.
Kendall B, Creaser RA, and Selby D (2006) Re-Os geochronology of postglacial black shales in Australia: Constraints on the timing of “Sturtian” glaciation. Geology 34: 729–732.
Nebel O (2014) Rb–Sr dating. In: Rink W and Thompson J (eds.) Encyclopedia of Scientific Dating Methods. Dordrecht: Springer.
Patterson CC (1956) Age of meteorites and the Earth. Geochimica et Cosmochimica Acta 10: 230–237.
Shen JJ, Papanastassiou DA, and Wasserburg GJ (1996) Precise Re–Os determinations and systematics of iron meteorites. Geochimica et Cosmochimica Acta 60: 2887–2900.
Stein HJ, Markey RJ, Morgan JW, Hannah JL, and Schersten A (2001) The remarkable Re-Os chronometer in molybdenite: How and why it works. Terra Nova 13: 479–486.
Vervoort JD (2014) Lu-Hf dating: The Lu-Hf isotope system. In: Rink W and Thompson J (eds.) Encyclopedia of Scientific Dating Methods. Dordrecht: Springer.
Vervoort JD, White WM, Thorpe RI, and Franklin JM (1993) Postmagmatic thermal activity in the Abitibi Greenstone Belt, Noranda and Matagami districts: Evidence from whole rock Pb
isotope data. Economic Geology 88: 1598–1614.
Walker RJ, McDonough WF, Honesto J, Chabot NL, McCoy TJ, et al. (2008) Modeling fractional crystallization of group IVB iron meteorites. Geochimica et Cosmochimica Acta
72: 2198–2216.
White WM (2015) Isotope Geochemistry. Wiley.

You might also like