Chapter 2

Download as pdf or txt
Download as pdf or txt
You are on page 1of 44

The early Earth Chapter 2

The Solar System is thought to have begun after one or more local supernova
explosions about 4.6 Ga ago. In one widely accepted scenario of the later stages
of formation of the Solar System it is thought that there were hundreds of
planetesimals in the region occupied by the inner planets. Once the planetesimals
began to attain the proportions approaching those of planetary embryos it is likely
that the heat generated by collisions would have been sufficient to allow both
melting and, as denser materials began to sink inwards, separation of the original
constituents. However, the development of a more evolved layering, such as that
seen on Earth, would require this separation to have been more or less complete
once giant impacts had ceased. This is because from that point onwards further
evolution of the Earth would be mainly driven by processes from within the planet.
The first 660 million years of the Earth’s existence, known as the Hadean, was
the stage during which the metallic core separated from the silicate mantle, the
atmosphere and hydrosphere were formed, and melting of the silicate mantle
produced the earliest crust. The early Earth was violent and hot, and giant
impacts would have been devastating. The heat released would have been
capable of melting the outer part of the Earth to form a global magma ocean.
Convective currents and degassing would have also destroyed any solidified
surface areas. Consequently, there is virtually no direct rock record of the
Hadean. This means that we have to turn to theoretical modelling and
geochemistry to reveal the mechanisms of formation of the different layers in the
Earth and the timescales involved.
In this chapter, you will be introduced to the evidence for the development of the
early Earth. The first part reviews the various heat sources necessary to drive
planetary differentiation. The second part investigates the mechanisms of core
formation, the evidence for the presence of a magma ocean, and the timing of
accretion and core formation. The third part explores the evidence for the origin
and age of the Moon. In the fourth part, you will learn about the formation and
evolution of the atmosphere and hydrosphere, and finally, the evidence for the
nature and formation of the earliest continental crust will be reviewed. You will
discover that the processes that formed the layering in the early Earth,
atmosphere and hydrosphere effectively shaped the planet as we know it today,
and ultimately provided conditions suitable for life to develop. You will also see
that water played a key role, not only in the formation of the atmosphere and
hydrosphere, but also in the early development of the solid Earth.

2.1 Heating and differentiation of the Earth


Differentiation is the process by which planets develop concentric layering, with
zones that differ in their chemical and mineralogical compositions. The generation
of such zones results from a differential mobility of elements due to differences in
their physical and chemical properties.
One obvious way of mobilising constituents in a planetary body is by allowing
them to melt. When a rock is heated, different minerals within the rock will melt

51
Book 1 Our Dynamic Planet

at different temperatures. This phenomenon is known as partial melting and is a


key process in the formation of liquid rock or magma, as you will see in Chapter 4.
Once elements have been mobilised in this manner, they will begin to migrate
under the influence of pressure or gravity.

 Imagine a body the size of a planetary embryo that had accreted from
nickel–iron and silicate minerals. Nickel–iron has a density of about
7.9 × 103 kg m–3 (compared to ~3.0 × 103 kg m–3 for silicate minerals) and a
melting point some hundreds of degrees higher than silicates. What would
happen if temperatures within this planetary embryo were increased to a
point at which silicates began to melt?
 Since nickel–iron has a higher melting point, it would remain solid after the
silicates had begun to melt and, because it is much denser than any silicate
minerals, it would begin to sink towards the centre of the body.

In addition to the separation of a metallic core from a silicate mantle, partial


melting of planetary bodies also produces residual solid silicate minerals in
contact with a silicate melt that have different compositions, depending on how
extensive the melting is. In such a system, incompatible elements are partitioned
into the melt more readily than compatible elements (Box 1.3). This process,
known as element partitioning, is the principal mechanism by which incompatible
elements first become concentrated into the melt. Then, if the magma is buoyant,
they migrate upwards to form the overlying crust. If element partitioning
continues over time, this concentration and migration will gradually create a crust
and mantle with different compositions.

2.1.1 Heat sources


If partial melting is the principal cause of differentiation, then the Earth needs to
be heated before layering can begin to develop. There are several sources of
heat that can arise during Earth’s evolution. The most important are:
• the so-called primordial heat sources, which develop in the early stages of
planetary evolution (i.e. those associated with accretion, collision and core
formation)
• tidal and radiogenic heating processes, which can operate long after the
planet has been formed.
Accretional heating
During accretion, any planetesimal falling towards the Earth will acquire a
velocity because of gravitational attraction from the Earth. The body will thus
have a kinetic energy due to its motion. If the velocity of the body immediately
before impact is v, the kinetic energy (E) at that point will be:

1 2
E= mv (2.1)
2

52
Chapter 2 The early Earth

where m is the mass of the body. Upon hitting the Earth, if all the kinetic energy
of motion is converted into heat, then the increase in temperature, ∆T, can be
calculated:

mv2
∆T = (2.2)
2(m + M )C
where the body (of mass m) impacts the Earth (of mass M) and C is the specific
heat capacity of Earth material (i.e. the amount of heat required to raise the
temperature of 1 kg of material through 1 K).

Question 2.1
(a) A planetesimal of mass 1015 kg impacts the Earth with a velocity of
10 000 m s–1. Calculate the rise in temperature in the Earth assuming that the
heat generated by the impact spreads rapidly and uniformly throughout the
whole Earth. Because m is much smaller than M, the effect of m is negligible
and can be ignored, so Equation 2.2 can be simplified to:

mv 2
∆T =
2MC
Take the total mass of the Earth to be 6 × 1024 kg, and the average specific
heat capacity of the Earth to be 750 J kg–1 K–1. (Note: 1 J = 1 kg m2 s–2)
(b) Suppose that the Earth was constructed entirely of 1015 kg planetesimals,
each of which generated the temperature rise obtained in part (a). What
would be the total temperature rise?

In practice, the various uncertainties involved make it difficult to determine how


much heating occurred during accretion. For one thing, as you will see later in
this chapter, not all of the impacting material arrived at the same time: accretion
took place over ~107 years. More importantly, not all of the kinetic energy would
be converted to heat. For example, some of the impact energy would be used in
the excavation of large craters. Finally, much of the heat would have been
radiated into space. Nevertheless, most estimates predict temperatures to have
risen above the melting point of silicate minerals and iron–nickel, which means
that the Earth is likely to have gone through an early molten stage.

Core formation
If the Earth went through an early molten phase, allowing the metals and silicates
to separate, then the ‘falling inwards’ of the nickel–iron-rich fraction to form the
core would have released potential energy.

 What do you think would have happened to this gravitational potential energy?
 The gravitational energy lost by the inward movement of nickel–iron would
have been converted first to kinetic energy and then into thermal energy. It is
estimated that the core-forming process would have contributed significantly
to the Earth’s primordial heating (though it would still have been an order of
magnitude less than that generated by collision and accretion).

53
Book 1 Our Dynamic Planet

As you will see later in this chapter, geochemical evidence suggests that accretion
and core formation were completed very early in the Earth’s history but, rather
than being a single ‘catastrophic event’, it is likely to have been a gradual process,
with progressive segregation of the core as more material was accreted.
Since both accretion and core formation relate to events that occurred early in
the history of planetary evolution, they are primordial processes and the heat
generated by these processes was primordial heat. However, if these primordial
heat sources had remained the only way of heating the Earth, their intensity
would have waned through time due to continual radiative heat loss to space.
Since heat drives fundamental processes such as volcanism, the fact that the
Earth has remained volcanically active to the present day requires additional
processes of internal heat generation.
Tidal heating
One heat source known to be generated within planetary bodies is tidal heating,
which is created by the distortion of shape resulting from mutual gravitational
attraction. These effects are readily observed upon the Earth’s oceans where the
attraction created by the Sun and Moon produces ‘bulges’ in the ocean water
masses that are then dragged around the planet as the Earth rotates. This
process produces the ebb and flow of tides seen around the coast. In much the
same way, the solid Earth is also distorted by these forces and produces tides
that reach a maximum amplitude of about 1 m on the rocky surface. This
deformation causes heating within the planet, though precisely where this heating
is concentrated depends upon the planet’s internal properties. In Earth’s case, it
is thought to occur largely within the crust and mantle.

Question 2.2
The current rate of heating generated within the Earth by tidal distortion is
estimated at 3.0 × 1019 J y–1. Given the mass of the Earth is approximately
6.0 × 1024 kg, determine the rate of tidal heating. Express your answer in W kg–1
(1 W = 1 J s–1) and to an appropriate number of significant figures.

Radiogenic heating
During the latter half of the 19th century, the eminent physicist Lord Kelvin
(1824–1907) attempted to determine the age of the Earth. He believed the Earth
had cooled slowly after its formation from a molten body and assumed the main
sources of energy were from primordial heat and tidal friction. Taking many
factors into account, including the mass of the Earth, the current rate of surface
heat loss, and the melting points of various rock types, he concluded that the
planet could not be much older than about 20–40 Ma.
This conclusion was contrary to the ideas of eminent geologists such as James
Hutton (1726–1797) and Charles Lyell (1797–1875), who had already argued that
immense spans of time were required to complete the changes produced by the
action of tectonic, erosional and depositional processes. It was also greatly at
odds with the then emerging theories of the evolution of life because scholars
such as Darwin (1809–1882) also argued for much longer periods based upon the

54
Chapter 2 The early Earth

evidence of biological speciation. As a consequence, a ‘heated’ controversy


continued for many years. Even though Kelvin’s calculations did not gain wide
acceptance with geologists he was a powerful scientific influence of the time,
and it was not until much later in the 1950s that accurate radiometric dating
experiments eventually proved him wrong. These experiments were conducted
on primordial lead in meteorites and demonstrated that the formation of the Earth
occurred about 4.6 Ga ago (See Section 2.2).
So why did Kelvin get the age of the Earth so wrong? The answer lies partly in
the decay of unstable isotopes of certain radioactive elements, the discovery of
which was not made until some decades after Kelvin’s initial calculations. John
Joly (1857–1933), an Irish physicist, was one of the first to suggest that
radioactive decay, leading to radiogenic heating, was an important heat source
within the Earth. It is now known that radioactive decay creates an important
independent source of heat within the Earth that supplements that remaining from
primordial sources (Box 2.1). This radiogenic heating is something that Kelvin
could not possibly have known about when making his calculations.

Box 2.1 Radioactive decay and heating


Most elements have different isotopes (i.e. atoms having the same number
of protons but different numbers of neutrons). Some of these isotopes are
unstable and decay to stable forms. For example, isotopes deficient in
protons decay by the transformation of a neutron into a proton and an
electron, which is expelled from the nucleus. During this process, known as
beta-decay, the mass of the nuclide does not change significantly. In
contrast, during alpha-decay, heavy atoms decay through the emission of an
α-particle, which consists of two protons and two neutrons (He2+). This
process reduces the overall mass of the nuclide. α- and β-particle collision
with adjacent nuclei during decay causes heating through the loss of kinetic
energy.
The rate of decay of a radioactive parent nuclide to form a stable daughter
product is proportional to the number of atoms, n, present at any time, t:

dn
= −λn (2.3)
dt
where λ is the decay constant characteristic of the radionuclide in question
(expressed in terms of reciprocal time). The decay constant states the
probability that a given atom of the radionuclide will decay within a stated time.
dn
The term is the rate of change of the number of parent atoms, and is
dt
negative because this rate decreases with time. Equation 2.3 can be
integrated from t = 0 to t, where the number of atoms present at time t = 0 is
n0:
n = n0e–λt (2.4)

55
Book 1 Our Dynamic Planet

A useful way of referring to the rate of decay of a radionuclide is by its


half-life (t½), which is the time required for half of the parent atoms to decay.
n0
On substituting n = and t = t½ into Equation 2.4:
2
ln 2 0.693
t½ = = , where ln 2 is the natural log of 2. (2.5)
λ λ
The number of radiogenic daughter atoms formed (D*) is equal to the
number of parent atoms consumed (Figure 2.1). So:
D* = n0 – n (2.6)
The total number of daughter atoms (D) consists of those produced by
radioactive decay after time t (i.e. D*) plus those already present at time
t = 0 (i.e. D0), that is:
D = D0 + D*. (2.7)
As D* = n0 – n (Equation 2.6) we can substitute D* in Equation 2.7:
D = D0 + n0 – n
Since n0 = neλ t (from Equation 2.4):
D = D0 + neλ t – n
D = D0 + n(eλ t – 1) (2.8)

Figure 2.1 The changing


number of parent and
increasing daughter atoms (D*) daughter atoms during
number of parent and

radioactive decay. This


daughter isotopes

illustrates that the growth of


radiogenic daughter atoms
(D*) is the mirror image of
the decay curve of the
number of parent atoms
decreasing parent atoms (P)
(n0). The half-life (t½) is the
time taken for half of the
0 t1/2 time, t parent atoms to decay (at
which point n0 = D*).

56
Chapter 2 The early Earth

Table 2.1 lists a number of isotopes that either occur in the Earth or for which
there is evidence of their having been active at some time during Earth history.

Table 2.1 Naturally occurring radioactive decay systems of geochemical and cosmochemical interest.

Parent Decay Decay Half-life/y Present heat Daughter Measured


mode* constant (λ)/ y−1 production‡ isotopes ratio
40K β+, e.c., 5.54 × 10−10 1.28 × 109 2.8 40Ar, 40Ca 40Ar/36Ar

β–
87Rb β– 1.42 × 10−11 4.88 × 1010 87Sr 87Sr/86Sr
147Sm α† 6.54 × 10−12 1.06 × 1011 143Nd 143Nd/144Nd
187Re β 1.59 × 10−11 4.35 × 1010 187Os 187Os/188Os
232Th α 4.95 × 10−11 1.39 × 1010 1.04 208Pb, 4He 208Pb/204Pb,
3He/4He
235U α 9.85 × 10−10 7.07 × 108 0.04 207Pb, 4He 207Pb/204Pb,
3He/4He
238U α 1.55 × 10−10 4.47 × 109 0.96 206Pb, 4He 206Pb/204Pb,
3He/4He
26Al β– 9.5 × 10−7 0.73 × 106 26Mg 26Mg/24Mg
129I β– 4.41 × 10−8 1.57 × 107 129Xe 129Xe/130Xe
146Sm α 6.73 × 10−9 1.03 × 108 142Nd 142Nd/144Nd
182Hf β– 7.78 × 10−8 8.9 × 106 182W 182W/184W
244Pu α, SF 8.45 × 10−9 82 × 106 nXe nXe/130Xe**

*α = alpha decay (4He);β– = beta decay (electron or positron); e.c. is electron capture; SF is spontaneous fission.
†The production of 4He from 147Sm decay is insignificant compared with that produced by decay of U and Th.
‡Heat production averaged for the whole Earth in units of 10−12 W kg −1 of Earth material (not of the isotope).
**n can be 124, 126, 128 or 129, all of which are produced by 244Pu fission. Element symbols are listed in the Appendix.

 Which of the isotopes listed in Table 2.1 remain active today and which are
extinct?

 All those with half-lives significantly less than the age of the Earth, i.e. 4.6 Ga,
are extinct, namely: 26Al, 129I, 146Sm, 182Hf and 244Pu. The others, principally
isotopes of 40K, 87Rb, 147Sm, 232Th, 235U and 238U, are still active today.
If radionuclides have short half-lives and are not replenished by the decay of other
isotopes, then they may be lost altogether. One such short-lived extinct nuclide
is 26Al, which has a half-life of 0.73 Ma.

 What evidence would you look for to support the presence of 26Al in the early
Solar System?
 26Al decays to 26Mg (Table 2.1), so you would expect to see anomalously high
abundances of 26Mg relative to other isotopes of Mg in materials from the
early Solar System.

Studies of primitive meteorites, notably carbonaceous chondrites, do indeed show


slightly high 26Mg/24Mg ratios, which suggests that a significant proportion of the
aluminium present at the time of condensation of the solar nebula was the unstable
isotope 26Al. This was originally created during a supernova explosion predating

57
Book 1 Our Dynamic Planet

the birth of our Solar System, and cannot be replenished by the spontaneous
decay of other radiogenic elements.

 What does the observation that 26Al was present in chondritic meteorites tell
you about the timescale of the formation of the Solar System?
 The half-life of 26Al is only 0.73 Ma, so the time between the supernova
explosion that generated the 26Al and the accretion of the meteorite parent
body must have happened on a similar timescale of a few million years.

1
Given that after 10 half-lives only 2−10 (or th) of the original number of 26Al
1024
atoms remain, then for any measurable amount of radiogenic 26Mg to be found,
chondritic meteorites must have formed within, at most, 7.3 Ma of the supernova.
Similar arguments are used later in this chapter for the timescales of formation of
other planetary bodies and layers within the Earth. The decay of 26Al may have
contributed significantly to the heating of planetary embryos but, because of its
short half-life, any remaining 26Al has long since vanished within the terrestrial
planets.
Whilst such short-lived isotopes may have been important heat sources during the
early stages of terrestrial planet evolution, study of Earth material indicates that it
is the isotopes of the elements U, Th and K that are responsible for most of the
radiogenic heating that has occurred throughout the history of the planet. These
isotopes, which all have particularly long half-lives (see Table 2.1), are termed
long-lived radiogenic nuclides and were present in sufficient quantities after
condensation and accretion to ensure that they have remained abundant within
present-day Earth.
Most common minerals contain small amounts of the elements with these
unstable isotopes, the most important of which are summarised in the upper half
of Table 2.1. Their decay to form more stable isotopes releases tiny increments
of heat, as described in Box 2.1. This decay has produced a continuous source of
heat within our planet since the Earth’s formation. Of course, the total amount of
heat produced will depend upon the concentration and types of radiogenic isotope
present in the parts of the Earth, and the mass of suitable material present in its
different layers.

 U, Th and K are incompatible elements (i.e. they are preferentially


partitioned into silicate liquids), so where will they be concentrated in the
Earth?
 The elements U, Th and K (and their radiogenic isotopes) are particularly
concentrated in the silicate-dominated outer layers of the Earth, and in
particular within the continental crust. They are thought to be virtually absent
from the core.

As a result of the incompatibility of the heat producing elements, the radiogenic


heat produced per unit mass of the continental crust is, on average, over one
hundred times greater than that of the underlying mantle. But because the mantle
is so much more massive than the crust, in effect this means that the overall
radiogenic heating budget is roughly split equally between the mantle and crust
58
Chapter 2 The early Earth

despite the much greater mass of mantle material. It is the decay of these long-
lived isotopes that provides sufficient heat energy to keep the Earth geologically
active. Therefore, the surface heat flux is not simply the slow cooling of a once
molten body, as originally envisaged by Kelvin.
Finally, whilst the rate of radiogenic decay is constant for each isotope system
(Table 2.1), the total amount of radioactive decay, and hence heat generation, will
decline over time as the reserves of the original radioactive materials are gradually
used up. This gradual depletion of radioactive materials is expressed in half-lives,
and each isotopic decay system has its own unique half-life. For example 235U
decays through a series of α-particle emissions to the daughter isotope 207Pb. The
data in Table 2.1 indicate that, after 7.07 × 108 years, half of the 235U originally
present will have decayed to 207Pb and the remainder will continue to halve every
7.07 × 108 years. Over the age of the Earth (4.6 Ga), approximately 6.5 half lives
of 235U have elapsed, so the heat production from 235U is now (½)6.5 ≈ 0.01 (i.e.
1%) of what it was originally.

Question 2.3
(a) What proportions of the original 40K and 232Th currently remain since the
formation of the Earth?
(b) Based upon the data in Table 2.1, what was the amount of radiogenic heating
in the Earth at 4.6 Ga, and how does it compare with that of today?
(c) What proportion of Earth’s surface heat flux loss is due to radioactive decay,
compared with the 1.5 × 10–13 W kg–1 (the value determined in Question 2.2)
created by tidal heating effects?

2.1.2 Heat transfer within the Earth


In the previous section you have seen how accretion, core formation and
radioactive decay heated the Earth. But how is this internal heat transferred to the
surface? Three main mechanisms of heat transfer operate within the Earth; these
are conduction, convection and advection.
Conduction This is perhaps the most familiar mechanism, since it is the process of
heat transfer experienced when the handle of a pan becomes hot. Heat is conducted
from the stove to the pan and then to its handle. Different materials, such as rocks of
various compositions, conduct heat at different rates, and the efficiency of heat
transfer in this manner is known as conductivity. This method of heat transfer is the
most important in the outermost layer of the Earth (i.e. the lithosphere).
Convection This involves the movement of hot material from regions that are
hotter to those that are cooler and the return of cool material to warmer regions.
During this transfer the material gives up its heat. It is a particularly efficient
method of heat transfer, but the medium through which transfer takes place must
be fluid. It should be noted, however, that the term ‘fluid’ describes any substance
capable of flowing and is not just restricted to liquids and gases. Under the right
conditions, even solid rocks can flow, albeit at a very slow rate. Over long periods
of geological time, the effects of such solid-state flow become a highly significant
way of transferring heat towards a planet’s surface.

59
Book 1 Our Dynamic Planet

Advection The final process of transferring heat is when molten material (magma)
moves up through fractures in the lithosphere and remains there. This is termed
advection and operates when magma spreads out at the surface as a lava flow or,
if it is injected, cools and crystallises within the lithosphere itself. The effect is the
same in both cases, since heat is transferred by the molten rock from deeper levels
where melting is taking place to shallower levels where it solidifies, losing its heat
by conduction into the overlying crust. Any planetary body that exhibits, or has
exhibited, volcanic activity must have lost some of its internal heat in this manner.
In Chapter 1 you learned that under the conditions prevailing deep within the Earth
the solid rocks of the mantle can flow when subject to surface loads, leading to
isostatic readjustment of surface elevations. The mantle can also flow when
subject to temperature differences in a process known as solid-state convection
and, whilst rates may be no more than a few centimetres per year, it is the most
efficient form of heat transfer within all but the outermost part of the mantle.
Near the Earth’s surface the rocks are too cold and rigid to permit convection, so
conduction is the most significant process. You should recall from Chapter 1 that this
zone of the uppermost mantle and all of the overlying crust is called the lithosphere.
In the underlying asthenosphere the principal process of heat transfer is convection
and this process of heat transfer applies to much of the mantle thickness down to the
core (Figure 2.2a). The marked difference in strength between the lithosphere and
asthenosphere is caused by increasing temperature with depth below the Earth’s
surface (the geotherm) (Figure 2.2b). So the thickness of the lithosphere depends
on the rate at which temperature increases with depth.
Finally in this section, it is of interest to return to Kelvin and the debate concerning
the age of the Earth. Kelvin assumed that the Earth cools by conduction alone.
You can see from Figure 2.2b that the geotherm for the convecting mantle is much
shallower than that in the lithosphere. In this way, convection in the asthenosphere
maintains a higher geothermal gradient within the lithosphere than would occur by
conduction alone, thus mimicking the geotherm of a planet that is much younger

continental crust temperature/ ¡C


subduction zone 0 500 1000 1500
spreading ridge 0
ocean crust
elastic

lithosphere

lithosphere mechanical
boundary
upper mantle layer
100
asthenosphere thermal
(convecting mantle)
depth/km

lower mantle boundary


layer
asthenosphere

200
outer core
convecting
interior
inner core
300
(a) (b)

Figure 2.2 (a) Section through the Earth showing the division of the mantle into the uppermost rigid lithosphere and the
mobile, convecting asthenosphere. (b) Geotherm through the lithosphere and uppermost asthenosphere.
60
Chapter 2 The early Earth

than the Earth. By assuming that the surface, conductive geotherm applied to the
whole Earth, Kelvin arrived at an erroneously young age for the Earth – a result he
would have obtained even if he had included the correct estimates for radiogenic
heat production!

2.2 The age of the Earth and its layers


Throughout this and the preceding chapter the age of the Earth has been given as
being around 4.6 Ga. But where is the evidence for this? To find out just how old
the Earth is we once again have to return to meteorites and radioactivity, for, in
addition to being sources of heat in planetary systems, radioactivity also allows
absolute ages to be determined from measurements of long-lived radioactive
isotopes and their daughters.
Several isotope systems are used to date events and processes from throughout Earth
history, but the three most commonly used are the K–Ar, U–Th–Pb and Rb–Sr
systems. The principles of radiometric dating are most clearly illustrated using the
Rb–Sr system, as outlined in Box 2.2, and isotope data from this and the U–Th–Pb
system are most frequently illustrated on an isochron plot (or isochron diagram or
isotope evolution diagram), examples of which are shown in Figure 2.3.
Figure 2.3a shows Rb–Sr isotope data from a series of ordinary chondrites that
define an isochron age of ~4.50 Ga. This age relates to the last time the Rb and Sr
were fractionated from each other by a particular process. In the case of Rb and
Sr, both elements are lithophile (Box 1.3), so it is unlikely that they were fractionated
by the separation of a metallic phase from a silicate fraction. However, Rb, being a
Group 1 alkali metal, is significantly more volatile than Sr, a Group 2 element similar
to Ca, which is one of the early condensing elements (Figure 1.16). Hence the Rb/Sr
fractionation may relate to the loss of a volatile phase; the age indicates when the Rb/
Sr ratio in ordinary chondrites was last disturbed.
Figure 2.3b shows a slightly more complex plot of data relating to the U–Pb
system. You should notice from Table 2.1 that the U–Pb system has two parent

0.82 40
LL-chondrites age 4.55 ± 0.07 Ga
0.80
30
0.78
207Pb/204Pb

0.76
87Sr/86Sr

chondrites
20
0.74 carbonaceous
chondrites Ð selection
of chondrules
0.72 age 4.493 ± 0.018 Ga
10 sulfides
(87Sr/86Sri) 0.69882 ±0.00008
0.70 terrestrial leads
(Pacific sediments)
0.68
0 0.5 1 1.5 2 0 10 20 30 40 50 60
87Rb/86Sr 206Pb/204Pb
(a) (b)

Figure 2.3 (a) Rb–Sr isochron plot for a suite of ordinary chondrites, giving an age of 4.49 ±0.02 Ga. (b) Isochron plot
of 206Pb/204Pb against 207Pb/204Pb for chondritic meteorites giving an age of 4.55 ±0.07Ga. (Minster and Allègre, 1980;
Murty and Patterson, 1962)
61
Book 1 Our Dynamic Planet

isotopes, 235U and 238U, decaying to 207Pb and 206Pb respectively. By combining
these two it is possible to eliminate the U/Pb ratio and determine an age from the
plot of 207Pb/204Pb against 206Pb/204Pb. In this case, the ages represent the time at
which U was fractionated from Pb and, as Pb is a moderately volatile element and
can be lithophile, siderophile or chalcophile in different environments, it is less easy
to define the process that led to U/Pb fractionation. However, iron meteorites are
rich in Pb and poor in lithophile U, so the age probably represents the timing of the
separation of a metallic phase. Given that the chondrite isochron passes through
the Pb isotope ratio of most iron meteorites, it adds further support to this idea.
The data illustrated in Figure 2.3 are for whole meteorite samples but, as we have
seen in Chapter 1, meteorites are far from homogeneous, comprising a number of
different components. Primitive carbonaceous chondrites are thought to be
amongst the least differentiated material in the Solar System. Among other things,
they contain chondrules and calcium- and aluminium-rich inclusions (CAIs).
Chondrules are millimetre-sized spherical droplets believed to have been produced
when mineral grain assemblages were flash heated and cooled quickly. CAIs are
typically centimetre-sized and consist of the first minerals to condense at
equilibrium from a gas of solar composition. A detailed study of CAIs and
chondrules yielded a 206Pb/207Pb isotope age for CAIs is 4567.2 ±0.6 Ma, whereas
that of chondrules is 4564.0 ±1.2 Ma.

 What is the difference between the ages of CAIs and chondrules, and how old
then are carbonaceous chondrites?
 The data give an interval of 3.2 ±1.8 Ma between formation of the CAIs and
chondrules – carbonaceous chondrites must have formed at or after the time of
formation of the chondrules i.e. 4564 Ma.

Even though the difference between these two ages is small, it is greater than the
combined uncertainty associated with the two ages – they are significantly
different. The difference represents a real difference in the timing of the formation
of the CAIs and chondrules.
These data show that the oldest components of meteorites, and hence the Solar
System, must be close to 4.57 Ga old, but how do we know that this age also
applies to the Earth?
Part of the answer to this question lies in Figure 2.3b, where the average Pb
isotope ratios of Pacific sediments are compared with the data from chondrules.
The sediment data fall on or close to the meteorite isochron, implying ultimate
derivation from a similar source or common parent. Other evidence is found in
lunar samples returned by the Apollo missions, which have ages that extend back
through the Hadean. The oldest igneous rocks from the Moon are samples of
anorthosite. One clast has been dated at 4.56 ±0.07 Ga using the 147Sm–143Nd
system, placing the formation of the Moon to within 70 Ma of the start of the Solar
System. But this in itself raises the question of what we are trying to date – what
do we mean by the age of the Earth? In particular, given that radiometric dating
systems date the time of element fractionation and therefore reflect the effects of
different processes during Earth accretion and differentiation, it is probably more
precise to consider the timing of major processes in the formation of the Earth.

62
Chapter 2 The early Earth

Hence the following sections focus on determining the timing of important events
in the history of the early Earth in relation to the age of the Solar System, which
is given as 4.57 Ga.

Box 2.2 Radioactivity applied to dating


The rubidium–strontium isotope system provides a good illustration of the
principles of isotope dating, and will be used here to demonstrate those
principles. The number of 87Sr daughter atoms produced by the decay of
87Rb in a rock or mineral since its formation t years ago is given by

substitution into the radioactive decay equation (Equation 2.8).


87Sr = 87Sri + 87Rb(eλ t – 1) (2.9)
where 87Sri is the number of 87Sr atoms initially present. It is, however,
difficult to measure precisely the absolute abundance of a given nuclide.
Mass spectrometers can, however, measure isotope ratios to very high
accuracy and precision, and so it is more convenient to work with isotope
ratios by dividing by the number of atoms of 86Sr (which is a stable isotope
and therefore remains constant with time).

⎛ 87 Sr ⎞ ⎛ 87 Sr ⎞ ⎛ 87 Rb ⎞
⎜ 86 ⎟ = ⎜ 86 ⎟ + ( e − 1) ⎜ 86 ⎟
λt (2.10)
⎝ Sr ⎠ P ⎝ Sr ⎠i ⎝ Sr ⎠
y c m x
The present-day Sr isotope ratio (87Sr/86Sr)P is measured by mass
spectrometry, and the 87Rb/86Sr ratio can be calculated from the measured
concentrations of Rb and Sr. If the initial ratio (87Sr/86Sr)i is known or can
be estimated then t can be determined, if it is assumed that the system has
been closed to Rb and Sr mobility from the time t to the present.
Most rocks are many millions of years old, in which case it is difficult to
estimate the initial Sr isotope ratio. However, examination of Equation 2.10
shows that it is equivalent to the equation for a straight line.
y = c + mx (2.11)
By plotting (87Sr/86Sr)P on the y-axis against 87Rb/86Sr on the x-axis, the
intercept c is then the initial ratio of the system (Figure 2.4). On such a
diagram, a suite of co-genetic rocks or minerals having the same age define
a line termed an isochron, and the diagram is called an isochron plot. The
slope of the isochron, m = eλ t – 1, yields the age of the rocks or minerals. If
one of the rocks or minerals is Rb-poor then this may yield the initial 87Sr/
86Sr ratio directly. Otherwise, the initial ratio is determined by extrapolating

back to the y-axis using a best-fit line through the available data points.
The isotope evolution of a suite of hypothetical minerals is shown in Figure
2.4. At the time of crystallisation of the minerals all four have the same
87Sr/86Sr ratio, and plot as points on a horizontal line (AB). After each of

these minerals becomes closed to exchange of Rb and Sr isotopes, evolution


begins. The sloping line (AC) represents the 87Sr/86Sr and 87Rb/86Sr ratios
measured today.

63
Book 1 Our Dynamic Planet

 Study Figure 2.4. What has happened after time t?


 87Rb has decayed to 87Sr over the period of time t to the present day.
This has had two effects: (i) it has reduced the amount of 87Rb present,
so the 87Rb/86Sr ratios have decreased slightly; (ii) it has increased the
amount of 87Sr present and so the 87Sr/86Sr ratios have increased.

 But why have the 87Sr/86Sr ratios increased by differing amounts along the
line AC?
 The amount of 87Sr produced by radioactive decay depends on the
amount of 87Rb present. However, Figure 2.4 uses isotope ratios; hence
the relative increase in 87Sr (the increase in 87Sr/86Sr) depends on the
relative amount of 87Rb (or the 87Rb/86Sr ratio). Thus, after a given
period of time the samples with the highest 87Rb/86Sr ratio will also be
those with the highest 87Sr/86Sr ratio.

The four samples shown on Figure 2.4 represent an idealised case. For an
isochron to yield a slope that reflects a true age then the following
assumptions must be valid.
1 All samples must be of the same age. If the samples
are from the same igneous intrusion or are minerals
C
from the same rock then this is a reasonable
t=t assumption. However, if the samples are from a large
area where the geology is poorly understood, then
they may be of a different origin and may not be the
same age and will not plot on a straight line.
Sr/ Sr
86

2 All samples must have the same initial 87Sr/86Sr ratio


87

when they formed, otherwise they will not plot on the


same line on the isochron diagram (AB on Figure
2.4) even if they are of the same age. This
A B assumption is likely to be valid for a suite of igneous
t=0
rocks that have crystallised from the same parental
melt, but it is less likely for sediments that are
87 86
Rb/ Sr composed of differing amounts of recycled material.
Figure 2.4 Isochron plot of 87Sr/86Sr against 3 The 87Sr/86Sr and 87Rb/86Sr ratios of the samples
87Rb/86Sr illustrating how four samples of the have only changed by the process of radioactive
same age but different Rb/Sr ratios evolve from decay – no Rb or Sr has been added to or lost from
a horizontal line (AB) at the time of their them between the time of formation and the present
formation, to plot on a straight line (AC) with a day. Such open-system behaviour might occur due
slope m = eλ t – 1 at the present time.
to the infiltration or loss of fluids, or diffusional
exchange between minerals after their crystallisation.

64
Chapter 2 The early Earth

2.2.1 Core formation and magma oceans


The formation of the metallic core is the biggest differentiation event that has
affected the Earth, resulting in a large-scale change in the distribution of
composition, density and heat production. One would think that such a
fundamental feature would be well understood, but the physical mechanism by
which metal separates from a silicate mantle and accumulates at the centre of
the Earth remains poorly understood.
One potential mechanism for
Fe–Ni metal separation or θ <60¡ θ <60¡
segregation is that the metal melt
melts and forms an θ
interconnected network.
Whether or not this happens
depends on a property known
crystal crystal
as the dihedral angle, θ
(Figure 2.5). The dihedral angle
(a) (b)
is that formed by the liquid in
contact with two solid grains,
which in the case of the mantle will be silicate or oxide grains. If the dihedral Figure 2.5 The definition of the
angle θ is <60°, the melt will fill channels between the solid grains and form an dihedral angle θ, and the
interconnecting network, even in small melt fractions. On the other hand, if θ is depiction of two different
>60°, the melt is confined to pockets at grain corners and cannot easily move, microstructures for static
systems: (a) wetting, where melt
unless the melt fraction is greater than 10%.
forms an interconnected network
If melt is able to connect, its rate of migration is quite rapid, and can be along grain edges, and (b) non-
calculated using Darcy’s law: wetting, where melt forms
isolated pockets at grain corners.
k (Rushmer et al., 2000)
v= ∆ρ g (2.12)
η
where v is the velocity of the melt relative to the solid matrix, k is the
permeability, η is the viscosity of the melt measured in Pa s, ∆ρ is the density
difference between the melt and the solid, and g is the acceleration due to
gravity.
Permeability can be defined as:

a 2Φ
k= (2.13)
24π
where a is the mean grain radius and Φ is the melt fraction.

Question 2.4
Taking a grain radius, a, of 10–3 m (1 mm), Φ of 0.1 (10% volume melt), ∆ρ of
3500 kg m–3, g of 9.8 m s–2 (the acceleration due to gravity on Earth) and a
viscosity, η , of 0.005 Pa s, calculate the migration velocity of Fe–Ni metal (give
your answer in kilometres per year). (Note: 1 Pa s = 1 kg m–1 s–1)

Such calculations show that any metallic melt that can form an interconnected
network ought to sink rapidly to form a core. The key question then is the extent
to which the metal connects.
65
Book 1 Our Dynamic Planet

 Experiments indicate that the dihedral angleθ is >60° for metallic melts. To
what extent will such melts form an interconnected network?
 If θ is >60° then melts will be isolated at grain corners, creating an impermeable
silicate framework through which metallic melts cannot segregate.

For this reason core formation is thought by many to occur only after the silicate
framework has broken down after extensive silicate melting (>40%). At these high
degrees of melting the grain boundary framework will no longer be interlocked, but
rather crystals will be floating in a silicate liquid – a crystal mush. In such a mush,
dense molten metal droplets would sink, but to achieve such high degrees of
melting requires enormous amounts of heat.

 How might such heating have occurred on the early Earth?


 As you have previously seen, accretional heating during planetesimal–embryo
Earth collisions would result in extensive melting, which is likely to have
resulted in a global magma ocean.

It is important to note that there is no independent evidence that a magma ocean


ever existed on Earth. Any early formed crust has long since been destroyed by
impacts, erosion and plate recycling. Nevertheless, a deep magma ocean is thought
to be a likely consequence if accretion included giant impacts as predicted in
modern theory. The extent and depth of any magma ocean depends on many
factors, including the impactor/Earth mass ratio, impact velocity and initial
temperatures of the impactors.
The evidence also suggests that the Earth had a huge proto-atmosphere, formed
by degassing of the Earth’s interior. This would have provided a thermal blanket
that retained the heat generated during accretion and sustained the magma ocean.
Finally, as you will see later in Section 2.3, a giant collision is the most popular
theory for the origin of the Moon, and such an impact would have delivered
sufficient energy to melt the entire planet. These accretional and impact sources of
heat would also have been supplemented by radioactive decay, which would have
been much greater than at the present day.
If the entire silicate mantle were molten, the metal blobs would sink directly to the
centre of the Earth, but if the base of the magma ocean were solid silicate then
liquid metal might pool at this boundary until gravitational instability permits
movement of the metal to the core in large diapirs. Alternatively, metal from the
impactor and perhaps even the Earth’s proto-core may have become highly
fragmented and emulsified in the magma ocean, in which case small metallic
droplets could equilibrate rapidly with silicate melt.
Cooling of a peridotite magma ocean would eventually lead to crystallisation from
the bottom up. In the absence of an atmosphere, heat radiation to space would
have been very efficient and the deeper part of the magma ocean would have
cooled very quickly (i.e. in less than 1000 years). However, the upper part of the
magma ocean could have remained hot and molten much longer, maybe for more
than 107–108 years, especially if a cooled crust formed at the surface. This
situation may have been complicated by the continuous supply of new material
from meteorites that remelted on impact, and tidal heating.

66
Chapter 2 The early Earth

2.2.2 Core–mantle equilibration


You should recall from Chapter 1 that one consequence of core–mantle
separation is that the metal-loving siderophile elements would be strongly
partitioned into the metallic core. However, trace amounts of siderophile
elements are retained in the mantle and if metal segregation were an equilibrium
process then these elements would provide important clues for deducing the
conditions of core formation.
Figure 2.6 shows the abundances of siderophile elements subdivided into slightly,
moderately and strongly siderophile, in the Earth’s mantle normalised to
chondritic meteorites, revealing the stepped abundance profile as discussed in
Section 1.4. An assumption of many early core formation models was that metal
segregation was contemporaneous with accretion, and that metal and silicate
equilibrated at near-surface, low-temperature (T) and low-pressure (P)
conditions. Low T–P metal silicate distribution coefficients (Box 1.3) for highly
siderophile elements have been determined experimentally and found to lie
between 10–7 and 10–15. These values lead to the expected abundances in the
mantle as shown in Figure 2.6.

 How do the abundances of the highly siderophile elements observed in the


silicate mantle compare with what would be expected from the
experimentally determined low-pressure partition coefficients?
 The observed data (triangles in Figure 2.6) indicate a depletion of about
2 × 10–3 relative to chondrites (i.e. present at ~0.2% of the chondritic
abundance), whereas the partition coefficients predict depletions of 10–5 and
10–6 (filled circles in Figure 2.6).

As you have seen, metal segregation in a magma ocean would probably occur
over a range of temperatures and pressures, and so equilibrium metal segregation
at high temperature and high pressure in a deep magma ocean becomes a
realistic possibility. Applying experimentally determined partition coefficients to

10
slightly siderophile moderately siderophile highly siderophile
mantle normalised to Mg and chondrites

10Ð1

10Ð2 Figure 2.6 The abundance of


siderophile elements in the
10Ð3 present-day Earth’s upper
mantle. Also shown are the
10Ð4 mantle abundances expected for
selected elements in the mantle,
10Ð5
assuming separation of the core
as metal droplets sinking through
10Ð6
Mn V Cr Ga P Fe W Co Ni Sb As Ge Mo Os Re Ir Ru Pt Rh Au Pd a magma ocean and equilibrating
according to low- and high-
measured abundances predicted abundances predicted abundances pressure partition coefficients
in the mantle assuming low-pressure assuming high-pressure
D-values D-values respectively. (Modified after
Drake and Righter, 2002)

67
Book 1 Our Dynamic Planet

metal segregation at high pressures can reproduce the abundance of Re, but as
yet experimental data on other elements are lacking. However, given their
dramatically different partitioning behaviour at low temperature and pressure it is
unlikely that there exists any set of conditions at which all siderophile partition
coefficients converged to a single value, which is required by the uniform
depletion of the most highly siderophile elements.
The failure of low-temperature, low-pressure metal/silicate equilibration models
to explain the siderophile excess inspired a number of alternative models,
including:
• partitioning between a sulfur-rich liquid metal and silicate
• inefficient core formation, where small amounts of metal or sulfide remain
behind in the mantle
• the heterogeneous accretion or ‘late veneer’ model in which core
formation effectively strips out all the siderophile elements from the mantle,
which are subsequently raised to the observed values by another process.
This is the most popular model.

 What process could have raised siderophile element abundances in the


silicate mantle following core formation?
 Continued accretion of meteoritic materials from new impacts of meteorites
with chondritic proportions of the siderophile elements.

In contrast to the highly siderophile elements, the abundances of moderately


siderophile elements would not have been significantly altered by the later
addition of meteoritic material because they would not have been so efficiently
stripped from the mantle by core formation. Potentially, moderately siderophile
elements can provide much more information on the conditions of core formation.
Of these elements, Ni and Co provide some key constraints because their
abundances in the mantle are accurately and precisely known, and their
partitioning behaviour has been studied over a wide range of conditions.

 With reference to Figure 2.6, estimate the Ni/Co ratio of Earth’s mantle at
the present day.
 Figure 2.6 shows that within uncertainty, Ni and Co are present in proportions
that are close to chondritic, i.e. both at ~0.1 × chondrite.

The chondritic ratio of Ni to Co in the mantle requires the ratio of the two
partition coefficients DNi/DCo to be about 1.1. Experiments show that an increase
in pressure and/or temperature causes both Ni and Co to become less
siderophile, but at different rates (Figure 2.7).

 Assuming a DNi/DCo ratio of about 1.1 is required to explain the Ni/Co ratio
of the Earth’s mantle, use the experimental data in Figure 2.7 to estimate the
pressure of metal–silicate equilibration.
 From Figure 2.7, a DNi/DCo ratio of 1.1 occurs at a pressure of about 28 GPa
equivalent to a depth of 900–1000 km, implying high temperature and
pressure metal–silicate equilibration and core segregation.

68
Chapter 2 The early Earth

Figure 2.8 summarises the conceptual model of metal


segregation and metal/silicate equilibrium in a deep magma 100
ocean. For the model geotherm shown, the entire upper mantle nickel
cobalt
would be molten (a magma ocean), whereas the lower mantle
would be solid.
In the upper part of the mantle, equilibrium metal segregation D
10
from the upper mantle would occur by the ‘rain-out’ of small,
liquid metal globules over a wide range of temperature and
pressure conditions, and the metal would accumulate at the
DNi/DCo = 1.1
magma ocean floor. At the boundary between the lower and
upper region, the metal would equilibrate a final time, giving the
upper mantle its present siderophile element signature. Finally, 1
0 5 10 15 20 25 30
gravitational instability would cause the formation of large metal pressure/GPa
diapirs that sink through the lower region, with or without re-
equilibration. (Note: these divisions do not necessarily Figure 2.7 Metal/silicate partition coefficients (D)
as a function of pressure for nickel (Ni) and cobalt
correspond to the present-day upper and lower mantle.)
(Co). See text for explanation. (Walter et al., 2000)

chilled crust Figure 2.8 Pressure–temperature


0 diagram showing metal/silicate
equilibrium and metal segregation
10 mantle iron droplets in a deep magma ocean.
solidus molten (a) Pressure–temperature plot of
20 mantle
Pressure/GPa

liquidus the mantle solidus and liquidus


iron ponds (Chapter 4) compared with the
30 mantle
solid mantle geotherm. Where the
diapir
40 model geotherm geotherm lies above the liquidus,
the mantle is completely molten
50 (upper region); where it lies
core below the solidus it is completely
60 solid (lower region). (b) Cartoon
1000 2000 3000 4000
section through the Earth’s mantle
(a) temperature/ ¡C (b) early in Earth history showing
metal droplets falling to the base
2.2.3 The timing of core formation of the mantle magma. The
droplets accumulate into larger
In order for any radioactive decay system to be of use in dating a process, that bodies of dense, molten metal that
process must be able to fractionate the parent element from the daughter eventually sink through the solid
element. Thus, in order to investigate the timing of core formation, radioactive lower mantle as diapirs and
systems are needed in which one of either the parent or the daughter elements is accumulate in the core. (Modified
siderophile and the other is lithophile. from Walter et al., 2000)

 Which of the radioactive decay schemes in Table 2.1 satisfy this criterion?
(Hint: use Figure 1.24 to determine the dominant geochemical properties of
the elements involved.)

 The two decay schemes with elements of contrasting properties are hafnium–
tungsten (Hf–W) and uranium–lead (U–Pb).

Of the two systems, U–Pb involves long-lived isotopes, whereas the Hf–W
scheme has a much shorter half-life – it is much easier to illustrate the principles
of the approach to this problem using the Hf–W system.

69
Book 1 Our Dynamic Planet

The hafnium–tungsten system


In addition to the contrasting geochemical properties of Hf and W during core
formation, both elements are refractory and so were accreted to the Earth in
chondritic proportions, therefore the Hf/W ratio of the bulk silicate Earth is
known relatively well.
 Assuming the bulk silicate Earth has a chondritic Hf/W ratio, how will the
Hf/W ratio of the core and mantle differ from that in chondrites?
 The core will have an Hf/W ratio lower than chondrites because W is
siderophile and will be enriched in the core. Hf, being lithophile, will remain in
the mantle, which will have an Hf/W ratio greater than in chondrites.
182Hf decays by β decay to 187W with a half-life of 8.9 Ma (Table 2.1).
 If the core separated from the mantle magma ocean while 182Hf was
sufficiently abundant, what would happen to the 182W/184W ratio of (a) the
mantle and (b) the core?
 At the time of element fractionation, the 182W/184W ratio of both core and
mantle would be similar. However, with time, that of the mantle would
increase rapidly because of its high Hf/W ratio, while that of the core would
increase less rapidly because of its low Hf/W ratio.

The effect of Hf/W fractionation on W isotopes is well illustrated by measurements


of W isotopes in iron meteorites. The data are illustrated in Figure 2.9 and reveal
that the metal from iron meteorites has
bulk silicate low 182W/184W ratios (low ε182Hf) (see
Earth Box 2.3), lower than both chondritic
meteorites and the silicate Earth (the
mantle). The simplest explanation for
this is that these metals sampled early
Martian Solar System tungsten before live 182Hf
meteorites had decayed.
the Moon

carbonaceous
chondrites

Figure 2.9 ε182W for iron meteorites,


carbonaceous chondrites, the bulk
silicate Earth and the Moon. These show
iron a well-defined deficiency in 182W in
meteorites early metals and carbonaceous
chondrites relative to the bulk silicate
Earth (BSE). Error bars denote analytical
uncertainty on individual measurements.
(Data from Foley et al., 2005; Klein et
Ð6 Ð4 Ð2 0 2 4 6 al., 2002; Yin et al., 2002; Scherstein et
182 W
al., 2006; Lee et al., 2002)

70
Chapter 2 The early Earth

Box 2.3 Tungsten isotope ratios and their notation


The beta decay of 182Hf to 182W (t = 9 Ma) has ⎛ 182 W ⎞
proven to be of enormous value in determining the in which ⎜ 184 W ⎟ is the measured 182W/184W
⎝ ⎠sample
relative timing of events during planetary accretion
and core formation. Because 182Hf is now extinct, ⎛ 182 W ⎞
evidence for its original presence is recorded in the ratio of the sample and ⎜ 184 W ⎟ is the
⎝ ⎠ BSE
isotope ratios of its daughter element, tungsten (W).
measured 182W/184W ratio of the bulk silicate Earth.
Tungsten consists of five stable isotopes but only one
of these, 182W, has been partly produced by the Consider the following example. A chondritic
radioactive decay of 182Hf. Therefore, to use this meteorite has a measured 182W/184W ratio of
geochronometer, geochemists need to measure the 0.864640. In the same experiment, a standard
abundance ratio of 182W to another of the isotopes representative of the bulk silicate Earth has a
of W, conventionally 184W. measured 182W/184W ratio of 0.864810.
The magnitude of the variations in the 182W/184W What is the ε182W value of the meteorite?
ratio relate to the parent/daughter ratio, 182Hf/184W,
in the original material. This was of the order of ⎡⎛ 0.864640 ⎞ ⎤
10–4. Consequently, variations in the 182W/184W ratio ε 182 W = ⎢⎜ ⎟ − 1⎥ ×10
4
⎣⎝ 0.864810 ⎠ ⎦
will be of a similar magnitude.
= (0.999803 − 1) ×104
Modern mass spectrometers are capable of
measuring isotope ratios to very high precision, in = −0.000196 ×104
some cases to better than 10 parts in a million = −1.96 = −2.0 (to 2 sig. figs)
(0.001%). This means that isotope ratios can be and
often are quoted to six decimal places. Dealing with This result shows that the chondrite measured has a
such numbers is inconvenient, especially when slightly lower amount of radiogenic 182W than the bulk
significant differences occur in the fourth and fifth silicate Earth by 0.02%, or 2 parts in 10 000. Quite
decimal places. To cope with this, geochemists have clearly, saying it has a ε182W value of −2 is much
adopted a convention in which differences in isotope easier than dealing with a six figure decimal, while the
ratios are compared to a standard value, usually one sign conveys that it is less radiogenic than BSE. A
that relates to the bulk silicate Earth or meteorites. sample with a positive value has more radiogenic W
In the case of tungsten isotopes, differences in than the BSE.
isotope ratios are measured in the number of parts in
104 (10 000) and are designated by the Greek letter The ε notation also has another advantage in that it
ε (epsilon), defined according to the following allows results from different laboratories to be
equation: compared more easily.

⎡ ⎛ 182 W ⎞ ⎤ Question 2.5


⎢⎜ 184 ⎟ ⎥ Calculate the ε182W value for the following data from
⎝ W ⎠sample
ε 182 W = ⎢⎢ 182 − 1⎥⎥ ×104 (2.14) another laboratory in which the same meteorite has a
⎛ W⎞
⎢ ⎜ 184 ⎟ ⎥ measured 182W/184W ratio of 0.864523 and a standard
⎢⎣ ⎝ W ⎠ BSE ⎥⎦ representative of the BSE has a measured 182W/184W
ratio of 0.864696.

71
Book 1 Our Dynamic Planet

The answers to the above calculations are given to 2 significant figures


because the precision with which individual measurements are made is of
the order of 0.3ε units, so any result is subject to this degree of uncertainty.
Both sets of original data give the same result within the limits of
experimental uncertainty, despite giving different absolute values for the
measured ratios. This illustrates that while different laboratories may get
different absolute values of 182W/184W ratios, even though they are
measuring the same samples, the calculation of ε182W values overcomes
these differences, as long as the samples and standards are measured in the
same laboratory.

But what about the Earth’s core? Unfortunately we do not have access to the
core but we can compare W isotope ratios in the silicate part of the Earth (the
mantle and crust) with chondritic meteorites – which are representative of the
bulk silicate Earth at least as far as Hf and W are concerned. The results in
Figure 2.9 show that the bulk silicate Earth has a different ε182W from
chondritic meteorites.

 Is the difference in W isotopes between chondrites and the bulk silicate


Earth consistent with core formation?
 Yes, the bulk silicate Earth has a higher ε182W and hence a higher
182W/184W ratio than chondrites, signifying that it has a higher Hf/W ratio.

The tungsten isotope difference between early metals, carbonaceous


chondrites and the bulk silicate Earth reflects the Hf/W ratio of the material
that formed the Earth and its fractionation during the lifetime of 182Hf. As a
result of core formation, the bulk silicate Earth has an elevated Hf/W ratio
(~15) relative to chondrite meteorites (Hf/W ratio ~1). Therefore, provided
that the Earth’s core formed early, i.e. during the lifetime of 182Hf, an excess
in the 182W atomic abundance in the bulk silicate Earth relative to chondrites
or iron meteorites will be generated. The conclusion from the extinct Hf–W
system is that planetary differentiation occurred while 182Hf was still present.
As its half-life is 8.9 Ma, this implies that the core and mantle must have
separated within the first few tens of millions of years of Earth history.
Needless to say, the detailed interpretation of W isotopes in planetary bodies is
more complex than described here, but the principles remain the same. The
data for chondrite meteorites, some of which are independently dated, yield an
ε182W value of around –2, relative to the bulk silicate Earth (and the Moon),
which has a much higher Hf/W ratio. These differences have been modelled
with core separation between 30 Ma and 50 Ma after the start of the Solar
System.

72
Chapter 2 The early Earth

As a footnote to this section, studies of Pb isotopes in the Earth and meteorites


have also been exploited to investigate the timescales of core segregation, again
based on the contrasting properties of U and Pb. The U–Pb system has a long
half-life and so any measurements of Pb isotopes in modern Earth materials
reflect a long history of U–Pb fractionation. However, it can be argued that the
major U–Pb fractionation occurred early in Earth history, that it was related to
core formation, and that it occurred over a period of 70–150 Ma after the start
of the Solar System.

2.3 The Moon


The Moon is a significant part of the terrestrial system and no consideration of
the history of the early Earth would be complete without at least a brief
consideration of its formation. One of the fundamental discoveries of the Apollo
missions was that the Moon consists of a variety of igneous rock types that
differ widely in their mineralogy, composition and age. The most visible
evidence of these differences is the existence of two distinct terrains on the
Moon:
• the light-coloured feldspathic rocks of the highlands
• the dark basalts of the maria (see Figure 1.2b).
Both rock types are igneous and indicate that at some time during its history the
Moon was extensively molten. Indeed the anorthosites, which are also very old,
are thought to have formed by a process of floatation on a magma ocean, their
dominant mineral, plagioclase, being lighter than the iron-rich basalts from which
they crystallised.
As discussed previously, the oldest igneous rocks are generally considered to be
the iron-rich anorthosites, which may be as old as 4.56 ±0.07 Ga. Since
anorthosites are igneous rocks, they provide a lower limit for the age of the
Moon, but again imply that the Moon must have formed within the first 70 Ma
of the start of the Solar System.
It is also possible to use Hf–W model ages to date the age of the Moon. One
lunar basalt yields a positive ε182W, i.e. it has more radiogenic W than the bulk
silicate Earth (Figure 2.9). However, the Moon is also thought to possess a
higher Hf/W ratio than Earth (see below), so the derived age is similar to that of
the bulk silicate Earth, and ages for Hf/W fractionation in the Moon range from
between 30 Ma and 45 Ma after the start of the Solar System.
Structurally, the Moon contrasts with the Earth in that it has a very small core
and is totally solid – there is no convecting outer core and hence no magnetic
field. The Moon, being much smaller than the Earth and having a much larger
surface to volume ratio, cooled very rapidly after formation. As a result, the
lithosphere is very thick, comprising in excess of 75% of the thickness of the
Moon’s mantle.

73
Book 1 Our Dynamic Planet

The compositions of lunar rocks have been used to develop models of the bulk
composition of the Moon, which can be compared with that of the bulk silicate
Earth and meteorites. These are summarised for a selection of elements with
contrasting properties in Table 2.2. (Element names and symbols are listed in the
Appendix.)

Table 2.2 Comparison of elemental abundances in primitive meteorites, Earth and Moon.

CI chondrite Earth (crust Moon (crust Ratio of trace


(primitive + mantle) + mantle) element abundance
meteorite) Moon/Earth
Volatile elements
K/ppm 545 180 83 0.46
Rb/ppm 2.32 0.55 0.28 0.51
Cs/ppb 279 18 12 0.67
Moderately volatile
Mn/ppm 1500 1000 1200 1.20
Refractory elements
Cr/ppm 3975 3000 4200 1.40
Th/ppb 30 80 112 1.40
Eu/ppb 87 131 210 1.60
La/ppb 367 551 900 1.63
Sr/ppm 7.26 17.8 30 1.69
U/ppb 12 18 33 1.83
Siderophile elements
Ni/ppm 16500 2000 400 0.200
Mo/ppb 1380 59 1.4 0.024
Ir/ppb 710 3 0.01 0.003
Ge/ppm 48 1.2 0.0035 0.003
Note: some elements are parts per million (ppm) and some are in parts per billion (ppb).

 Relative to the Earth, what do you notice about the abundances of the volatile
elements in the Moon?
 They are all lower than those of the Earth.
One of the primary observations of the Moon is that it is depleted in the most
volatile elements and enriched in refractory elements. This has been interpreted
as relating to a very high temperature origin for the material that makes up the
Moon.

 How do the siderophile element abundances in the Moon compare with those
of the bulk silicate Earth?
 They are much lower.

74
Chapter 2 The early Earth

Explaining this difference is more complex, especially as the Moon may not
possess a metallic core. The extreme depletion of the siderophile elements in the
silicate portion of the Moon strongly suggests that the material of which the Moon
is made was already differentiated – it had already lost a metallic fraction and
hence its inventory of siderophile elements.
In addition to these characteristics, any model of lunar formation must also take
into account the following:
• the angular momentum of the Earth–Moon system. Angular momentum is a
property of rotating systems that depends upon mass and its distribution,
angular velocity, and radius. The angular momentum of the Earth–Moon
system is contained in the Earth’s rotation and the Moon’s orbital motion, and is
unusually high compared with the other terrestrial planets
• the Earth and Moon have indistinguishable oxygen isotope compositions,
whereas most planetary bodies have different and distinct oxygen isotope
compositions.

2.3.1 The formation of the Moon


The origin of the Moon has been debated for over a century, but particularly since
the Apollo mission provided samples to study. Several theories of formation have
been suggested.
Co-accretion
This theory proposes that the Earth and Moon simply accreted side by side. The
difficulty with this model is that it does not explain the angular momentum of the
Earth–Moon system. This model explains neither the difference in density nor the
difference in the depletion of volatile elements.
Capture
This theory proposes that the Moon was originally in a heliocentric orbit and was
captured following a close approach to the Earth. However, it is difficult to do this
without the Moon spiralling into the Earth and colliding. It is also difficult to
explain the indistinguishable oxygen isotope compositions of the Earth and Moon
with this model.
Fission
This theory proposes that the Moon split off as a blob during the rapid rotation of a
molten Earth. George Howard Darwin (1845–1912), the son of Charles Darwin,
originally championed this idea. At one time (before the young age of the ocean
floor was known) it was thought by some that the Pacific Ocean might have
been the residual space vacated by the loss of material. This model does have
some attractive features: it explains why the Earth and Moon have identical oxygen
isotope compositions. It also explains, for example, why the Moon has a lower
density, because the outer part of the Earth would be deficient in iron due
to core formation, and it explains why so much of the angular momentum of the
Earth–Moon system is in the Moon’s motion. However, it is not clear why the
Moon should spontaneously split away from the Earth without some large input
of energy.

75
Book 1 Our Dynamic Planet

Impact models
Following the Apollo missions it was proposed that the Moon formed as a result of
a major impact on the Earth that propelled sufficient debris into orbit to produce
the Moon. Such models are now the most widely accepted (Figure 2.10).
Important information that came from sample-return was that more than 80% of
the lunar crust was composed of anorthosite, indicating the presence of a magma
ocean early in lunar history. The presence of a magma ocean requires significant
degrees of melting, which occurred in response to an impact – provided the
accreting body was large and that subsequent lunar accretion was rapid (1–100 Ma).
Also, it is necessary to link the dynamics of the Moon with that of the Earth’s spin.
This led to the proposal of a series of single giant impact models in which the
Moon was the product of a glancing blow collision with another differentiated
planet (Figure 2.10a–b). A ring of debris would have been produced from the outer
silicate portions of the Earth and the impactor (named Theia, the mother of Selene,
the goddess of the Moon), which was roughly the size of Mars (Figure 2.10c).
This model explains the angular momentum, the extensive early melting, the
isotopic similarities and the density difference. The most recent model simulations
indicate that the giant impact that formed the Moon probably occurred at the very
end of Earth’s accretion, when the Earth was 90% of its present size.
Certain features of the Moon may be a consequence of prior differentiation of
Earth and Theia, and of the giant impact itself. For example, the depleted volatile
elements require that both Earth and Theia had already become at least partially
differentiated due to earlier heating. In addition, the low abundances of siderophile
elements can be attributed to prior core separation in both the Earth and Theia.

 How is the sequence of events described above consistent with the


observation that the Moon’s core (if any exists) is very small relative to the
size of the whole Moon?
 The Moon formed from mantle material derived from the colliding, partially
differentiated planetary embryos. This mantle material was already depleted in
Ni, Fe, and S due to the development of cores within the embryos. Therefore,
there would have been relatively little Ni, Fe and S left to differentiate inwards
once the Moon had formed.

 Figure 2.10 The formation of the Moon. (a)–(c) illustrate the collision and aftermath between the proto-Earth and
Theia. Both bodies were large enough to have differentiated into core, mantle and primary crust as a result of
accretionary heating. Following collision (a and b), the cores of the two bodies are thought to have combined and the
mantles became mixed while some material was fragmented and vaporised and scattered into orbit around the Earth.
(c) Some of the debris fell back to Earth while the remainder accreted under its own gravity to form the Moon.
(d) The heat of accretion would again have resulted in wholesale lunar melting. The Moon then cooled and
differentiated into mantle, primary crust and possibly a small core, depending on how much of the core of Theia was
dispersed around the Earth and how much merged with the Earth’s core. (e) The Moon and, presumably, the Earth
were subject to further meteorite bombardment and the formation of large craters. Some of these impacts were intense
enough to initiate melting of the lunar mantle, flooding the larger impact structures with basalt and forming the lunar
maria. (f) Over time the Moon’s orbit has slowly decayed, its rate of rotation becoming synchronised with its orbit as
the distance between the two bodies progressively increases.

76
Chapter 2 The early Earth

(a)

(b)

(c)

(d)

(e)
(f)

77
Book 1 Our Dynamic Planet

For the Moon’s crust and mantle, this depletion was probably further augmented
by further differentiation immediately after its formation. Moreover, the Moon’s
depletion in volatile elements, relative to the Earth, can be explained if the Moon
accreted from the partially vaporised debris coalescing after the impact. In these
circumstances, the more volatile elements would have had the opportunity to
escape into space prior to accretion.

Question 2.6
Given the arguments regarding planetary accretion, volatile elements, and their
behaviour following giant impacts, suggest a reason why the concentrations of
Rb, K, and Na differ between chondrites and the Earth’s mantle.

2.3.2 The late heavy bombardment


The final major event that affected the early Moon is known as the late heavy
bombardment. The chronology of impact events on the Moon has been based on
telescopic observations, crater counting and disturbed radiogenic isotope
systematics. The majority of impact-melt rocks returned from the Moon yield
ages around 3.85 Ga (Figure 2.11). These ages have been interpreted as either
representing a short and intense heavy bombardment or as the tail end of a
prolonged post-accretionary bombardment. In either case the bulk of this
bombardment, which produced craters many hundreds of kilometres across,
preceded 3.8 Ga. Within this time span, the Earth must have been subjected to a
significantly greater bombardment than the Moon, as it has a larger diameter and
a much greater gravitational cross-section, making it an easier target to hit. It has
been estimated that the impact rate for the Earth would have been more than 20
times greater than the Moon,
with both more and larger
impact events. The
consequences for the
Copernicus hydrosphere, atmosphere and
Aristillus even the lithosphere must
Autolycus have been devastating. It has
Imbrium been suggested that the
absence of any rocks on
Serenitatis
Earth older than ~4.0 Ga is
Crisium the result of this late heavy
Nectaris bombardment, during which
impact-induced mixing
0 1 2 3 4
crater age/Ga recycled early crustal
fragments back into the upper
Figure 2.11 Crater ages on the Moon. mantle.

78
Chapter 2 The early Earth

2.4 The origins of the atmosphere and


hydrosphere
The atmosphere and hydrosphere represent the products of the outgassing of the
Earth over geological time, primarily from volcanic activity. Yet the present-day
composition of the atmosphere does not reflect that of volcanic volatiles. These
are dominated by water vapour, carbon dioxide (CO2) and sulfur dioxide (SO2)
with smaller amounts of nitrogen, halogens, hydrides and other more exotic
volatile compounds. Most of these volatile compounds are soluble in water or, in
the case of methane, are easily oxidised to water-soluble compounds. They
therefore do not accumulate in the atmosphere but dissolve in the oceans and
react with the oceanic crust. The remaining less-reactive and less-soluble
elements and compounds, notably nitrogen and the inert gases (He, Ne, Ar, Kr
and Xe), accumulate in the gaseous atmosphere.

 Which important atmospheric gas is not abundant in volcanic volatiles?


 Oxygen.
For about the first billion years of Earth’s history, oxygen was only present in the
atmosphere in trace amounts as a result of the breakdown of water vapour by
UV radiation. However, this inorganic mechanism of releasing oxygen into the
atmosphere produces only tiny amounts of free oxygen. Since the Earth’s present
atmosphere is oxygen rich and because all higher forms of life require free
oxygen, there is an obvious need for some other source of oxygen. The most
plausible source is oxygen-producing photosynthesis, a process that first evolved
in cyanobacteria during the Archaean era (from 3.8 Ga to 2.5 Ga ago). So, in
part, our currently breathable atmosphere is a by-product of life and not a
primary feature of the geosphere alone.
The evolution of the hydrosphere is also intimately linked with that of the
atmosphere – water is a volatile compound and is only present on the Earth’s
surface because the surface temperature is below 100 ºC. So to understand the
origins of the atmosphere and hydrosphere our attention needs to be focused on
those components of the present-day atmosphere that have a limited or negligible
interaction with the modern biosphere, namely nitrogen and the inert gases.
An important factor dictating whether or not an object in the Solar System can
retain an atmosphere is the strength of the gravitational field at its surface − the
stronger the field, the stronger the gravitational forces acting on the molecules in
the atmosphere. Without the gravitational field, those molecules moving away
from the planet would be lost. Even with the gravitational field, those molecules
with particularly high velocities can still escape. This leads to the notion of
escape velocity, which is the minimum velocity needed before a body has
enough kinetic energy to escape from the surface of a planet (i.e. overcome its
gravitational field). It can be shown that the escape velocity (Vesc) for a body of
mass M and radius r is given by:

2GM
Vesc = (2.15)
r

79
Book 1 Our Dynamic Planet

where G is the universal gravitational constant. Whether atmospheric molecules


have sufficient velocity depends on the temperature. As the temperature of a gas
increases, its molecules move around more quickly, and the average velocity of its
molecules increases. Some fraction of the molecules will always be travelling fast
enough to overcome gravitational forces, allowing them to escape to space. At low
temperatures, this proportion is negligible, but at higher temperatures it becomes
progressively more significant, until most molecules exceed the escape velocity for
the planetary body. Note that the relevant temperature is that at a level in the
upper atmosphere above which the atmosphere is so thin that a molecule moving
outwards has little chance of colliding with another, and so will escape if it has
sufficient velocity.
Different gases have different molecular masses, so their average velocities are
different at a given temperature. In order for a planetary body to retain a particular
gas in its atmosphere for a period of time of the same order as the age of the Solar
System, the average velocity of the molecules in the gas should be less than about
one-sixth of the escape velocity. (If the average velocity exceeds one-sixth of the
escape velocity, a significant proportion of molecules will be moving faster, and will
be lost.) This condition is achieved on only a few planets and satellites.
Figure 2.12 explores these relationships further. For each of the planets (and Pluto
and the Moon), the temperature is plotted along the horizontal axis and one-sixth of
the corresponding escape velocity on the vertical axis. Note that in order to cover
the range of values needed, the scales are not linear, a particular distance along an
axis corresponds to a doubling of the quantity. For the named gases, the sloping
lines show at each temperature the average molecular velocities of each gas.
Figure 2.12 thus defines the conditions under which a planet would lose or retain
that gas over geological timescales, i.e. thousands of millions of years. The giant
planets plot well above all the lines; they can therefore retain any of the gases.
The Moon plots below all the lines – it can retain no gases.

20
Jupiter
10
8 Saturn
one-sixth escape velocity/km s−1

Neptune
4
Uranus
Figure 2.12 Graph summarising
conditions of absolute 2 Venus
Earth
temperature (K) and escape
velocity for which planetary 1 Mars hydro
gen
bodies can retain common gases 0.8 helium
Mercury water
in their atmospheres for long vapou
r; amm
periods. For planetary bodies 0.4 onia
Moon nitrog ; methane
with substantial atmospheres, the en; ox
ygen
temperature is that at the top of 0.2 Pluto carbo
n diox
the atmosphere. For other bodies ide
it is the mean surface 0.1
temperature. Pluto’s temperature
varies greatly according to its 1000 800 400 200 100 50
temperature at the top
position around its orbit. Note:
of the atmosphere/ K
both axes are logarithmic scales.

80
Chapter 2 The early Earth

 Which gases should Earth be able to retain?


 Earth plots below the hydrogen and helium lines so it cannot retain these
gases. It plots above the lines for water vapour, nitrogen, ammonia, methane,
oxygen and carbon dioxide, so it can retain these gases.

2.4.1 Nitrogen and the inert gases


Nitrogen is the most abundant gas (78.1%) in the atmosphere; the third most
abundant is argon (0.93%), one of the inert gases. The remaining inert gases, Ne,
Kr, and Xe are also found in small traces in the atmosphere. All are relatively
unreactive elements in inorganic systems and, although nitrogen can be removed
from the atmosphere (fixed) by bacterial activity, its presence in the atmosphere
in relatively large quantities is a reflection of its inorganic chemical inertness.
Much attention has been paid to the isotopic composition of the inert gases both
in the atmosphere and in the mantle because they can tell us much about the
sources of the Earth’s volatile elements. In particular, the isotopes of Ar and Xe
are significant because they are the daughter products of radioactive decay
schemes (Table 2.1), which can be exploited to tell us about the timescales of
planetary degassing. The reason why Ar is one of the more abundant
atmospheric gases is that it is dominated by one isotope, 40Ar. 40Ar is the
daughter of radioactive 40K, which has a half-life of 1.28 Ga (Table 2.1). Xe is
much less abundant than Ar, but one of its isotopes, 129Xe, is the daughter of 129I,
but in this case the radioactive parent has a half-life of only 15.7 Ma.

 What information can these two radioactive systems provide?


 By analogy with short-lived isotopes in the solid Earth (e.g. Hf–W) and in
meteorites (Mg–Al), the I–Xe couple provides information on the timescales
of outgassing in the early phases of Earth evolution. By contrast, the long
half-life of 40K means that the isotope ratio of Ar reflects the history of
planetary outgassing over the whole age of the Earth.

How does this work? The principles behind these isotopic systems are described
in Box 2.4. This explains how the radiogenic isotope ratios of the atmosphere and
the mantle evolve with time and why measurements of the isotope ratios of Xe
and Ar in both are necessary to understand global outgassing.

Box 2.4 Ar and Xe isotopes and the evolution of the


atmosphere
The effects of planetary outgassing on the K–Ar and I–Xe radioactive
systems are shown graphically in Figure 2.13, in which global outgassing is
regarded to be a single event that occurred early in Earth history. Prior to
this event, the mantle had low I/Xe and K/Ar ratios and so the ratio of the
daughter isotope to a non-radiogenic isotope increased steadily. After
degassing, the two parent–daughter ratios in the mantle increased markedly,
because Xe and Ar have been lost to the atmosphere and the isotope ratios
of Xe and Ar in the mantle both increased at a greater rate than before.

81
Book 1 Our Dynamic Planet

outgassing
event

mantle
129Xe/130Xe

atmosphere

arbitrary units
40Ar/36Ar

129Xe/130Xe
mantle

(a) time (b) t1 time

Figure 2.13 Schematic diagrams showing the effect of planetary degassing on


the isotopic evolution of Xe and Ar in the Earth’s atmosphere and mantle. (a) The
simple situation in which the atmosphere gradually forms from the mantle by
degassing. Early in Earth history there is a period when the 129Xe/130Xe ratio
evolves rapidly, because of the decay of 129I, until it reaches a maximum value
when all the 129I has decayed. Because 129I has a half life of 15.7 Ma this occurs
between 100 and 150 Ma after the formation of the Solar System. (b) If the
atmosphere formed as the result of a major outgassing event, t1, early in Earth
history while 129I was active, then the fractionation of Xe into the atmosphere
and the preferential retention of I in the mantle would result in greater 129Xe
production in the mantle. See text for further discussion.

The critical issue concerns the timing of the degassing event relative to the
half-life of 129I. The example illustrated in Figure 2.13b shows how the isotope
ratio of 129Xe/130Xe varies if degassing occurs before all the 129I has decayed.
Subsequent to degassing, the mantle and the atmosphere evolve along
different paths.

 What will be the consequences for the 129Xe/130Xe ratios of the mantle
and atmosphere if degassing occurs after all the 129I has decayed?
 They will be identical.
Note that the important measurement to make is not the Ar and Xe isotope
ratios of the atmosphere, but those of the mantle. The reason for this is that any
outgassing event or process extracts virtually all of the Xe or Ar from the
mantle and so the radioactive clock is reset. Because all the volatile elements
eventually end up in the atmosphere, its isotope composition reflects the sum
total of all outgassing events and processes throughout Earth history. Hence the
measurement of Xe isotopes in the Earth’s mantle provides a means of
distinguishing between early, catastrophic planetary outgassing as opposed to
gradual outgassing over the whole of Earth history.

82
Chapter 2 The early Earth

As you may imagine, making isotope measurements of Ar and Xe in the mantle is


extremely difficult and enormous care needs to be taken to avoid atmospheric
contamination, both of the samples and in the measuring instruments. However,
the most successful measurements show that both the 129Xe/130Xe and the
40Ar/36Ar ratios of the mantle are high relative to the atmosphere.

This observation requires early and extensive degassing of the mantle to generate
high parent/daughter ratios in the mantle early enough to have an effect on the
I–Xe system. Taken together, the Xe and Ar isotope ratios of the present-day
mantle show that between 80% and 85% of the atmosphere was outgassed
extremely early in Earth history, and given that the half-life of 129I is only 15.7 Ma
this suggests that outgassing occurred within the first few tens of millions of
years after accretion.

 Which major events occurred within the first few tens of millions of years of
Earth evolution?
 Core formation and the giant impact that formed the Moon both occurred
within 30–50 Ma of Earth accretion.

The short timescale of these two major events and the development of the
earliest atmosphere were all part of the primary differentiation of the Earth.
However, Earth history is never that simple – some outgassing continues to the
present day, largely via mid-ocean ridge and within-plate volcanic activity, but the
Xe and Ar isotope record of the atmosphere and the mantle show that little if any
of the noble gases and, by analogy, nitrogen, are recycled via plate tectonics or
any other process into the mantle.

2.4.2 Loss of the earliest atmosphere


Xenon isotope data also provide evidence that much (>99%) of the Earth’s early
atmosphere was lost within the first 100 Ma. The present I/Xe ratio of the Earth
is an order of magnitude higher than chondritic values at the start of the Solar
System. We know the abundance of 129I at that time, and as all of this 129I
formed 129Xe, it should have produced xenon that was highly enriched in 129Xe,
given the Earth’s high I/Xe ratio. However, instead, the Earth has xenon that is
only slightly more radiogenic than is found in meteorites that are rich in primordial
gases. This provides evidence that the Earth had a low I/Xe ratio that kept its
xenon isotope compositions close to the chondritic value. At some point xenon
was lost, but by this time 129I was close to being extinct so that the xenon did not
become radiogenic despite a very high I/Xe ratio.
We have already seen that gravitational forces are likely to prevent many gases
from escaping Earth (Figure 2.12), so it is difficult to see how a large early
atmosphere would have been lost. One possibility is that the atmosphere was
blown off by a major impact like the Moon-forming giant impact. Alternatively,
hydrogen and helium may have been present in substantial amounts in the early
atmosphere but, as Figure 2.12 shows, were easily lost through thermal escape.
If large amounts of these gases escaped rapidly, then other heavier gases could
be lost at the same time by being carried along with them.

83
Book 1 Our Dynamic Planet

Thus, the noble gas data indicate that >80% of the atmosphere was formed
extremely early in Earth history, and that much of this early atmosphere was lost
in the first 100 million years of Earth history. This suggests that initially the
atmosphere was far denser and more massive than today’s atmosphere. The
presence of such an atmosphere would have an insulating effect, leading to the
build up of temperatures in the outer parts of the Earth, facilitating the formation
of magma oceans and core formation.

2.4.3 Water
The evolution of the hydrosphere is intimately linked with that of the atmosphere
because water is also a volatile compound. Until recently, there were competing
views as to the origin of the Earth’s water. One was that the Earth accreted as a
dry body and its water was subsequently added through cometary impact. The
alternative view was that the Earth inherited its water from water-bearing
minerals in the undegassed interiors of planetary embryos, and that this was
outgassed, along with Xe and Ar, early in Earth history. As noted earlier, the
evidence suggests that the early Earth experienced intense meteoritic
bombardment and must have been hot. It might be assumed, therefore, that
surface conditions were too extreme for the young Earth to have a liquid
hydrosphere and that much of the initial water was lost to space. However,
current models suggest that with the presence of a dense early atmosphere the
pressures at the surface of the Earth beneath this atmosphere would have been
high enough to ensure that a significant proportion of water and other volatiles
were retained in solution. This debate was partially resolved with the
measurement of the deuterium/hydrogen (D/H) ratio in three comets, using both
space probe measurements (the Giotto probe to comet Halley) and two ground-
based measurements of radio and infra-red emissions. All three measurements
agree within experimental uncertainty and show that deuterium (heavy hydrogen)
is twice as abundant relative to hydrogen in comets as it is in the terrestrial
hydrosphere. Such a major distinction effectively rules out comets as a major
source of the Earth’s water.
Thus the preferred model for the evolution of the hydrosphere is that it degassed
from the mantle, and that this material was ultimately derived from water-bearing
grains that became incorporated into planetesimals and eventually into planetary
embryos.

2.5 The earliest continents


2.5.1 Isotopic evidence
As stated earlier, evidence suggests that, in its early stages, Earth may have had a
magma ocean, sustained by heat from accretion and the blanketing effects of a
dense early atmosphere. With the loss of the early atmosphere during planetary
collisions, the Earth would have cooled quickly, the outer portions would have
solidified, and it would have developed its first crust.
We have little idea of what such a crust might have looked like. This is because,
unlike the Moon, the Earth appears to have no rock preserved that is more than
4.0 Ga old. There was intense bombardment of the Moon until about 3.85 Ga
84
Chapter 2 The early Earth

(Section 2.3) and Earth’s earliest crust may therefore have been destroyed by
such impacts. It may also be that because Earth was hotter this crust was highly
unstable. Some have argued that the earliest crust was like the lunar highlands –
made from a welded mush of crystals that had previously floated on a magma
ocean. Others have suggested that it was made of denser rocks more like the
basalts that presently form the Earth’s ocean floor.
Some of the oldest rocks on Earth come from exposures in western Greenland
near Isua. These include ancient sediments and volcanic lavas that have since
been subjected to folding and intrusion by younger igneous rocks.
Metamorphosed sediments from Isua give an age of about 3.9 Ga, which is
regarded as a minimum age of sediment deposition. However, these rocks were
formed by weathering and erosion of pre-existing crustal material, and may
preserve clues as to the age and origin of this older crust.
Just as U–Pb and Hf–W are ideal for studying the rates of accretion and core
formation, and I–Xe is useful for studying the rate of formation of the
atmosphere, the 146Sm–142Nd (samarium–neodymium) system is useful for
studying the early history of melting in the silicate Earth. 146Sm decays to 142Nd
with a half-life of 103 million years (Table 2.1), but unlike Hf–W or U–Pb
systems, both Sm and Nd are lithophile elements and remain in the mantle during
core formation. However, during melting of the silicate mantle to produce the
crust, Nd is more incompatible than Sm (that is, Nd is preferentially partitioned
into the silicate melt) producing a crust with a low Sm/Nd ratio and leaving the
mantle with a complementary high Sm/Nd ratio. If melting of the silicate Earth
occurred within the lifetime of the now extinct 146Sm (about 500 Ma), then this
would fractionate Sm/Nd and could potentially produce differences in the
abundance of 142Nd in early crust. If subsequent recycling of such crust by
weathering and erosion or meteorite impact failed to eradicate the difference in
142Nd, then this might be detectable in early Archaean rocks.

However, such 142Nd differences (sometimes termed anomalies), if they ever


existed, were always likely to be small because of the low abundance of 146Sm in
the early Solar System. As a consequence their measurement is extremely
difficult and for many years only one sample showed the hint of such an effect,
and these data were questioned by many. Study of 142Nd anomalies has, needless
to say, focused on the Isua rocks and improved measurement techniques have
recently confirmed the existence of differences in 142Nd (from the present-day
mantle) in many rock types seen at Isua, including metamorphosed sediments and
volcanic rocks (Figure 2.14). These results indicate that the precursor materials
of the Isua rocks were derived from material that differentiated from the silicate
mantle within 50 Ma of the start of the Solar System.
Surprisingly, it was also discovered that an excess of 142Nd is ubiquitous in all
modern terrestrial rocks – at least all those measured so far – relative to
chondritic meteorites (Figure 2.14). If it is assumed that chondritic meteorites are
representative of the material from which the Earth was made, then this excess
142Nd indicates that the silicate Earth must have experienced a global chemical

differentiation within 50 Ma of the start of the Solar System.

85
Book 1 Our Dynamic Planet

Figure 2.14 142Nd/144Nd ratios bulk silicate


measured for chondrites, Earth
present-day terrestrial rocks and
Archaean rocks at Isua.

metabasalts

Isua orthogneisses

metasediments

modern
ocean basalts

carbonaceous
chondrites

Ð0.6 Ð0.4 Ð0.2 0 0.2 0.4 0.6


142Nd

2.5.2 Geological evidence


The rocks themselves at Isua allow us to make some inferences about conditions
on and in Earth 3.8 Ga ago. They include lavas known as komatiites (Box 2.5) that
have what is called a pillow structure, which indicates cooling under water, and
sediments rich in quartz pebbles and volcanic debris (Figure 2.15). The observation
that sediments and lavas were deposited and erupted under water indicates that
there must have been bodies of liquid water at the Earth’s surface, and perhaps
even ocean basins by 3.9 Ga. Land areas would have to have been exposed to
weathering and erosion to form such sediments. The presence of quartz pebbles
implies that at least some of the land being weathered was similar to the upper
parts of present-day continental areas. For example, granite is a common rock that
occurs in continental areas and is rich in quartz. It seems, therefore, that the
geological processes and cycles that are recognised today were operating on the
Earth at least 3.9 Ga ago, albeit with some differences in detail.

 What does the occurrence of sedimentary rocks at Isua tell us about the
surface temperature of the Earth 3.9 Ga ago?
 For weathering, erosion and deposition of sediments to occur there must have
been both rain and liquid water present, implying surface temperatures of
between 273 K and 373 K (i.e. 0–100 °C).

86
Chapter 2 The early Earth

Figure 2.15 (a) Metamorphosed


pillow lava at Isua showing
pillows with dark rims in a
matrix of chert (fine-grained
silica). (b) Metamorphosed
conglomerate at Isua, showing
pebbles of quartz.

(b)

(a)

Box 2.5 Komatiites

Komatiites are rare ultramafic lavas that were produced most commonly
during the Archean and are very rare in the Phanerozoic. These magmas
are thought to provide a record of the thermal and chemical characteristics
of the mantle through time. Komatiites are distinguished by their high MgO
content (>18 wt%) and olivines showing a spinifex texture (elongated,
skeletal, branching crystals) (Figure 2.16), which were initially thought to be
indicative of the rapid quenching of magma. When komatiites were
subjected to melting experiments in the laboratory they were found to
possess very high liquidus temperatures – a result that was initially
interpreted as indicating deep melting at high temperatures. The progressive
decline of komatiites was then used as evidence for progressive cooling of
the Earth’s mantle.
Figure 2.16 Olivine
needles in a spinifex-
textured komatiite. Each
crystal is 2–3 cm long.

87
Book 1 Our Dynamic Planet

However, over recent years an alternative interpretation of komatiite


formation has gained favour. This is because komatiites are rich in water, and
experiments indicate that the liquidus temperature is dramatically reduced
under hydrous conditions. Similarly, in the presence of water spinifex textures
can develop even at a low cooling rate. Therefore it has been proposed that
the high water contents in komatiite lavas are due either to the Archaean
mantle containing a higher volatile content, left over from accretion, or that
water was introduced into the mantle as hydrated crust in a subduction zone
environment. This alternative interpretation of komatiites indicates that
water played a key role in their formation, and as a consequence the
Archaean mantle was only slightly hotter than the present day. It also
indicates that subduction may have operated in the early Archaean.

2.5.3 Continents and the hydrosphere during the


Hadean
The Archaean sediments and volcanic rocks at Isua and at other locations across
the world show that by about 3.9 Ga water was present on the surface of the
Earth, and that a cycle of erosion and deposition of continental rocks had been
established. Is there any evidence for water further back in time?
The possibly surprising answer to this question is ‘yes’, and the source of this
evidence is from a lonely outcrop of Archaean sandstones and conglomerates in
the Murchison district of Western Australia. These ancient rocks, which are
about 3 Ga old, are known as the Mount Narryer and Jack Hills quartzite units.
They contain detrital grains of the mineral zircon. Zircons are of enormous value
to Earth scientists because, once formed, they are almost indestructible and
contain high concentrations of U and Th and other important trace elements.
Hence they are one of the primary sources of high-precision ages from Pb
isotopes of crustal rocks and their trace element contents provide further
information on their formation conditions and provenance. Zircons form as
accessory minerals in igneous rocks, usually granites, but, because of their
indestructibility, they can survive the rock cycle of erosion, transport and
deposition and are frequently preserved in coarse-grained sedimentary rocks
such as sandstones, grits and conglomerates, and their metamorphic equivalents,
e.g. quartzite. Thus, the inescapable conclusion is that the Jack Hills zircons must
have been inherited from an older granite and so from older continental crust. So
far this is not very exciting – rocks are known that are almost 1 Ga older than the
Jack Hills quartzites, so why the excitement?
Detailed isotopic analysis of the Pb isotope composition of these zircons (Figure
2.17) has revealed that they range in ages from 3.05 Ga, which is close to the
age of deposition, up to much older ages – a significant number of crystals are in
excess of 4 Ga. Indeed, the oldest has an age of 4.4 Ga and formed only 170 Ma
after the start of the Solar System. These zircons are the oldest solid material yet
found on Earth and, because zircon can only form in high temperature igneous
processes in a high-silica magma, they are thought to represent a tiny sample
inherited from some of the Earth’s original continental crust.

88
Chapter 2 The early Earth

Figure 2.17 Images of a zircon from the Jack Hills quartzites in Western Australia:
(a) shows the damage produced during ion-microprobe analyses. Temperatures are
in degrees Celsius and refer to results from an assessment of their Ti contents. The
white ellipse shows the region of the crystal dated at 4.2 Ga; (b) shows the same
crystal from cathode-luminescence, revealing compositional banding. (Watson and
Harrison, 2005)

That conclusion is in itself exciting – solid material from the Hadean and
evidence for continental crust – but apart from their age, the Jack Hills zircons
show compositional variations that allow conditions of crystallisation to be
inferred. It turns out that the zircons crystallised at a temperature of around
700 ºC, a temperature that is no greater than the temperature that granites
crystallise today. As you will see in Chapter 6, granites can only form at this
temperature if water is present. They form by dehydration reactions in which
mica breaks down, releasing substantial amounts of water that dissolve in the
silicate melt so generated. The presence of granite and its crystallisation at low
temperatures is powerful evidence that water was present in the earliest crust of
the Earth.
This argument may appear a little convoluted, and the leap from the tiny Jack
Hills zircons to a planet with continents and an active low-temperature
hydrosphere is large, but the logic is sound. The information gleaned from these
zircons is that the surface environment during the Hadean may have been
remarkably similar to the Earth today!

2.5.4 The Archaean atmosphere, hydrosphere and


biosphere
The nature of the Earth’s atmosphere and hydrosphere after the Earth had
cooled and the first rocks were preserved at the surface is the subject of
considerable debate. It has been suggested that the warm conditions and
presence of water may have provided conditions suitable for life to develop.
Several geological settings might have hosted such early life. It may have been
supported in ocean basins during transient heating after a major meteorite impact,

89
Book 1 Our Dynamic Planet

or possibly life could have existed in a hydrothermal system. Submarine systems


would have also offered a protective setting against UV radiation.
Although the rocks at Isua show clear evidence that water was present at the
Earth’s surface, there are no fossils. Again, evidence for life comes from
geochemical clues. Carbonate in some of the volcanic rocks contains carbon with
a heavy isotope signature, i.e. it shows an enrichment in 13C over 12C. This is
complementary to the ubiquitous depletion in 13C relative to 12C, which is
characteristic of ‘light’ carbon that has been involved in biological processes.
However, the carbonate is all non-biological in origin and secondary (i.e.
precipitated after the crystallisation of the volcanic rock) and therefore may be
much younger than the host volcanic rock. Isotopically light carbon also occurs in
graphite flakes in some rocks and this chemical signature is thought by some to
be of biological origin. However, this interpretation remains controversial, as the
graphite is found in highly metamorphosed and deformed rocks, and is usually
associated with secondary carbonate deposition.
The sequence at Isua also contains rocks that are thought by some to provide
possible evidence for the changing nature of the Earth’s atmosphere as a result
of the appearance of life. Banded iron formations (BIFs) are characterised by
finely banded dark-brown, iron-rich layers alternating with lighter-coloured iron-
poor layers (Figure 2.18). The layers range in thickness from less than a
millimetre to about a centimetre. The iron-rich bands contain the highly insoluble
iron oxides: hematite (Fe2O3), limonite (Fe2O3.3H2O) and magnetite (Fe3O4).
Chert, a rock composed of precipitated silica, occupies the iron-poor bands.
Individual bands, often only a few millimetres thick, can extend for several
kilometres. How BIFs formed is not entirely clear, as we have no modern
analogues to guide us. However, the involvement of iron oxides in their formation
suggests that the process that led to the formation of BIFs must have affected
the oxidation state of iron and hence such rocks may contain information about
the oxidation state of the Earth’s ocean and atmosphere at that time.
Figure 2.18 Close-up view
of a banded iron formation.
The light bands are
dominated by silica-rich
chert and the darkest bands
by red–brown oxides of iron.
Each band is ~1 cm thick.

Whatever the chemistry occurring in the formation of BIFs, large amounts of


oxygen were incorporated into BIFs very early in the Earth’s history. One
interpretation is that this suggests oxygen was available in the shallow seas
where most BIFs were formed. Most theories for the origin of BIFs involve a
significant role for hydrothermal activity on the early Earth. The seawater

90
Chapter 2 The early Earth

flowing through these hydrothermal systems would have dissolved iron-containing


minerals, so that iron in a reduced form was subsequently injected into the deep
ocean through hydrothermal vents. It is generally accepted that the deep ocean
on the early Earth was extremely oxygen deficient or anoxic, so that iron
escaped oxidation and precipitation at the vents themselves but was deposited in
much shallower, more oxygenated water.
How did the iron get from the deep ocean to shallow water, crossing large
expanses of oceans in so doing? One idea is that the iron was actually consumed
by bacteria that flourished near the vents and that these bacteria then drifted
away in vast colonies into shallow water where they died, depositing a thin film
of organic-rich material. After a while the organic material would have been
recycled, leaving the iron behind in its highly insoluble oxide form.

 Hydrothermal vents provide a potential source for the large amounts of iron
involved in the formation of BIFs, but where might the large amounts of
oxygen come from?
 Various mechanisms have been proposed to account for the oxidation of BIFs
and the precipitation of Fe(III):
– abiotic breakdown of water by UV radiation
– direct oxidation by anaerobic bacteria
– photosynthesis, which generates free oxygen.

Summary of Chapter 2
The sequence of events accompanying the evolution of the early Earth is
summarised in Figure 2.19. Taken together, the evidence suggests that the
development of the core and magma ocean, and outgassing to form the
atmosphere, were contemporaneous and interdependent processes.
The issue of whether accretion of the Earth was homogeneous or heterogeneous
has long been debated. To some extent the Earth must have accreted
heterogeneously because early planetesimals are likely to have experienced
different degrees of differentiation. Furthermore, meteorites with different
chemical compositions from Earth continue to accrete right up to the present
day. However, if the early Earth were covered by a deep magma ocean, any
effect of accreting material with diverse compositions would be erased in the
magma ocean. A possible exception to this might be the highly siderophile
elements. The abundances of these elements in the silicate mantle might be
explained by metal–silicate equilibration in a magma ocean or may have been
added as a ‘late veneer’ after core formation.
Finally, it would appear inescapable that most of the water on Earth arrived as
part of the accretion process and the presence of that water influenced primary
differentiation. Accretion of hydrous materials and subsequent outgassing led to
the formation of a dense ‘steam’ atmosphere, which in turn served as a thermal
blanket allowing the long-term persistence of a terrestrial magma ocean. The
higher pressure of water in the atmosphere would have allowed the presence of

91
Book 1 Our Dynamic Planet

Figure 2.19 Schematic diagram time Ma major events


showing the timescale of major 4400 earliest granite
events in the evolution of the crust
early Earth. The timescale is mantle primary
crust
indicative rather than accurate. 4497 70 formation of
(Adapted from Abe et al., 2000) primary crust C M
degassing
end of Earth
4517 50 formation differentiation in
formation of magma ocean
the Moon
4527 40 chemical O A
differentiation
of protomantle gravitational loss by giant
separation impacts (?)
formation iron
4537 30
of core delivery of wet
planetesimals
dissolution
into silicate early loss
melt of atmosphere

4557 10 formation of
magma ocean

delivery of wet
planetesimals
capture of solar-type
formation of atmosphere (?)
proto-atmosphere

C core O magma ocean M mantle

initiation of
Earth formation A atmosphere and ocean
4567 0

liquid water at the Earth’s surface, and a significant amount of water would have
been able to dissolve in the silicate liquid of the magma ocean. Cooling and
crystallisation of the magma ocean would have provided a further source of
water for the atmosphere and hydrosphere, and a mantle reservoir of stable
hydrous phases.
The formation of the hydrosphere probably occurred soon after the atmosphere
because water is also a volatile molecule. The oldest zircons indicate that water
must have been present at the Earth’s surface 170 Ma after accretion. This is
because these ancient zircons crystallised from granite that formed at a
temperature only possible if water was present. It is also possible that hydrous
komatiites may have been produced either by an Archaean mantle with a high
volatile content, or by melting of hydrous crust in an early subduction zone
environment. Following the end of heavy bombardment of the Earth around
3.85 Ga, it is possible that early life may have been supported in ocean basins
heated by meteorite impact or hydrothermal systems.

92
Chapter 2 The early Earth

Learning outcomes for Chapter 2


You should now be able to demonstrate a knowledge and understanding of:
2.1 The different heat sources and transfer mechanisms involved in driving
planetary differentiation and the early chemical and physical evolution of
the Earth.
2.2 The current theories associated with the mechanisms and timing or
accretion, core formation, and the formation and existence of early magma
ocean.
2.3 The evidence for the origin and age of the Moon.
2.4 The evidence for the formation and chemical evolution of the Earth’s
atmosphere and hydrosphere.
2.5 The nature, composition and age of the earliest continental crust.
2.6 The role of partial melting and the presence of water in generating silicic
crustal material from the mantle.

93
94

You might also like