Chapter 2
Chapter 2
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.
51
Book 1 Our Dynamic Planet
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.
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?
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
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
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.
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.
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.
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?
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
lithosphere
lithosphere mechanical
boundary
upper mantle layer
100
asthenosphere thermal
(convecting mantle)
depth/km
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!
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.
⎛ 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
63
Book 1 Our Dynamic Planet
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
64
Chapter 2 The early Earth
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.
66
Chapter 2 The early Earth
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
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.
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
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
carbonaceous
chondrites
70
Chapter 2 The early Earth
71
Book 1 Our Dynamic Planet
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.
72
Chapter 2 The early Earth
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.
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.
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.
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.
78
Chapter 2 The early Earth
2GM
Vesc = (2.15)
r
79
Book 1 Our Dynamic Planet
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
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.
81
Book 1 Our Dynamic Planet
outgassing
event
mantle
129Xe/130Xe
atmosphere
arbitrary units
40Ar/36Ar
129Xe/130Xe
mantle
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
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.
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.
(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.
85
Book 1 Our Dynamic Planet
metabasalts
Isua orthogneisses
metasediments
modern
ocean basalts
carbonaceous
chondrites
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
(b)
(a)
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
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!
89
Book 1 Our Dynamic Planet
90
Chapter 2 The early Earth
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
4557 10 formation of
magma ocean
delivery of wet
planetesimals
capture of solar-type
formation of atmosphere (?)
proto-atmosphere
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
93
94