0% found this document useful (0 votes)
13 views

Chapter 3

Copyright
© © All Rights Reserved
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
13 views

Chapter 3

Copyright
© © All Rights Reserved
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
You are on page 1/ 22

Chapter 3: Seismicity

Seismic wave
Seismology is the study of earthquakes and seismic waves that move through and around
the earth. A seismologist is a scientist who studies earthquakes and seismic waves. Seismic
waves can be defined as, “the waves of energy caused by the sudden breaking of rock
within the earth or an explosion. They are the energy that travels through the earth and is
recorded on seismographs.

Types of Seismic Waves

There are several different kinds of seismic waves, and they all move in different ways. The
two main types of waves are body waves and surface waves. Body waves can travel
through the earth's inner layers, but surface waves can only move along the surface of the
planet like ripples on water. Earthquakes radiate seismic energy as both body and surface
waves. Recordings of seismic waves from earthquakes led to the discovery of the earth's
core and eventual maps of the layers of the Earth's inside. Just as the prism below refracts
light at its faces, seismic waves bend, reflect and change speed at the boundaries between
different materials below the Earth's surface.

Body Waves

Traveling through the interior of the earth, body waves arrive before the surface waves
emitted by an earthquake. These waves are of a higher frequency than surface waves.
P Waves

The first kind of body wave is the P wave or primary wave. This is the fastest kind of
seismic wave, and, consequently, the first to 'arrive' at a seismic station. The P wave can
move through solid rock and fluids, like water or the liquid layers of the earth. It pushes
and pulls the rock it moves through just like sound waves push and pull the air. Have you
ever heard a big clap of thunder and heard the windows rattle at the same time? The
windows rattle because the sound waves were pushing and pulling on the window glass
much like P waves push and pull on rock. Sometimes animals can hear the P waves of an
earthquake. Dogs, for instance, commonly begin barking hysterically just before an
earthquake 'hits' (or more specifically, before the surface waves arrive). Usually people can
only feel the bump and rattle of these waves.
A P wave travels through a medium by means of compression and dilation. Particles are
represented by cubes in this model.
P waves are also known as compressional waves, because of the pushing and pulling they
do. Subjected to a P wave, particles move in the same direction that the wave is moving in,
which is the direction that the energy is traveling in, and is sometimes called the 'direction
of wave propagation.
S - Waves
The second type of body wave is the S wave or secondary wave, which is the second wave
you feel in an earthquake. An S wave is slower than a P wave and can only move through
solid rock, not through any liquid medium. It is this property of S waves that led
seismologists to conclude that the Earth's outer core is a liquid.

S waves move rock particles up and down, or side-to-side perpendicular to the direction
that the wave is traveling in (the direction of wave propagation).
Surface Waves
Travelling only through the crust, surface waves are of a lower frequency than body waves,
and are easily distinguished on a seismogram as a result. Though they arrive after body
waves, it is surface waves that are almost entirely responsible for the damage and
destruction associated with earthquakes. This damage and the strength of the surface
waves are reduced in deeper earthquakes. Surface wave also classified as Love and
Raleigh waves.
A Rayleigh wave rolls along the ground just like a wave rolls across a lake or an ocean.
Because it rolls, it moves the ground up and down and side-to-side in the same direction
that the wave is moving. Most of the shaking felt from an earthquake is due to the Rayleigh
wave, which can be much larger than the other waves.
Love-waves are shear waves that causes back & forth (like a snake) ground movement.

Stress vs Strain Relationships


When rocks (or other solid materials) are subjected to differential (directional) stress, they
respond by deforming. Tensional stress stretches materials, compressional
stress squeezes them, and shear stress causes slippage and translation. The term used to
describe the deformation of materials is strain, which is defined as the change in size or
shape (or both) of a solid as a result of stress. Uniform stress causes a solid to change size
uniformly in all directions, whereas
differential stress results in a change in shape
and perhaps in size.
When a solid is subjected to increasing stress,
it passes through three stages of deformation
in succession:
1. Elastic deformation is a reversible (non-
permanent) change in volume of shape.
When the stress is removed, the solid returns
to its original shape and size. Sir Robert
Hooke (1635-1703) demonstrated that a plot of stress vs strain for material behaving in an
elastic fashion is a straight line, as shown in the figure to the right. There exists, however, a
limiting stress, known as the elastic limit (point Z), beyond which a solid suffers permanent
deformation and does not return to its original shape.
2. Ductile deformation is an irreversible change in size and/or shape in solids that have
been stressed beyond the elastic limit. If the stress is removed at point X', the material will
partially return to its original shape -- a permanent strain, equal to XY, has been introduced
into the material.
3. Fracture occurs in a solid when the limits of both elastic and ductile deformation are
exceeded. A material is said to be brittle when it deforms by fracture (breaking). Different
materials have different stress-strain relationships. Elastic materials such as rubber are
dominated by the elastic portion of the curve. Materials such as wood have some elastic
properties and virtually no region of ductile character. Common rocks exhibit elastic and
ductile behavior before ultimately breaking by brittle fracture.

The speed of propagation and the type of motion propagated depends upon the elastic
properties of the material, namely

1. density:
2. bulk modulus: (compressibility)
3. shear modulus: (twistability)

Elastic Moduli
The linear relationship between stress and strain is known as Hooke's Law and is specified
by five elastic moduli or elastic constants which express the ratio of a particular stress to a
resultant strain. Four important constants are explained briefly next.
Young's Modulus
Consider a rod (figure right) of length l, cross-sectional area A, to which a uniform force F is
applied to each end (a tensional force in this case):

Young's modulus, E, is given by;

Poisson's Ratio

The rod will also contract in radius (because it is being extended). The radial strain is
r/r.

Poisson's ratio is given by

Bulk modulus

Consider an initial volume V which is subjected to a hydrostatic pressure (force per unit
area is the same in each direction). Let V be the change in volume.

The bulk modulus is .


Shear modulus

If we apply a force parallel to a surface the force per unit area is known as the shear
stress . Consider a rectangular cube. Application of shear stresses to the top and bottom
interfaces produce a strain proportional to. The shear modulus is

3.3.4. Ray paths and wave fronts

Ray paths are radial lines drawn from the source it shows the direction in which the seismic
waves are propagating.

Figure -3.10 Ray paths and wave fronts

Wave fronts are surfaces along which portions of a propagating wave are in phase. For example,
arrival of the wave in figure 3.11 occurs were particles first move as the wave approaches (where
front at 00 phase). The maximum amplitude of the particle motion occurs along the 90 0 phase
wave front. Other wave fronts correspond to positions where the wave goes from positive to
negative amplitude (1800) and at the minimum amplitude (2700).
Figure 3.11 wave fronts are surfaces along which particle motions of the propagating wave are in
phase (one complete oscillation is 3600 of phase).
In a homogenous medium (constant seismic velocities), the body waves (P and S) radiate
outward along spherical wave fronts, while Rayleigh waves (R) rolling along the surface.
Variations in body wave’s velocity cause wave fronts to deviate from perfect spheres, thus
bending or refracting the ray paths. Ray paths can be used to analyze portions of seismic waves
that make it back to the surface. Wave fronts are perpendicular to ray paths.

Figure 3.12 a) initial wave fronts in homogeneous medium for compressional (P), shear (S), and
Rayleigh (R) waves. b) Wave fronts in heterogeneous medium for propagating P- wave.

3.3.5. Ray paths in layered medium


What happens when an incident compressional wave strikes a boundary between two media with
different velocities of wave propagation and/or different densities? Part of the energy is reflected
from the boundary and the rest is transmitted (refracted) into the next layer. At an interface
between two rock layers there is generally a change of propagation velocity resulting from the
difference in physical properties of the two layers. At such an interface, the energy within an
incident seismic pulse is partitioned into transmitted (refracted) and reflected pulses. The relative
amplitudes of the transmitted and reflected pulses depend on the velocities and densities of the
two layers, and the angle of incidence on the interface, and described by Zoeppritz-Knott
equations in terms of seismic velocities and densities of the two layers.

i) Reflection and refraction of normally incident seismic rays

Normal incidence rays are rays that hit the interface at right angle. Consider a compressional ray
of amplitude A0 normally incident on an interface between two media of differing velocity and
density (Fig. 3.13). A transmitted ray of amplitude A2 travels on through the interface in the same
direction as the incident ray, and a reflected ray of amplitude A1 returns back along the path of
the incident ray.

Figure- 3.13 Reflected and transmitted rays associated with a ray normally incident on an
interface of acoustic impedance contrast.

Assuming no loss of energy along any ray path, the total energy of the transmitted and reflected
rays must equal to the energy of the incident ray (i.e. A 0 = A1 + A2). The relative proportions of
energy transmitted and reflected are determined by the contrast in acoustic impedance (Z) across
the interface. The acoustic impedance of a rock is the product of its density (ρ) and its wave
velocity (v); that is,

Z= ρv

It is difficult to relate acoustic impedance to a tangible rock property but, in general, the harder a
rock, the higher is its acoustic impedance. Intuitively, the smaller the contrast in acoustic
impedance across a rock interface the greater is the proportion of energy transmitted through the
interface. Obviously all the energy is transmitted if the rock material is the same on both sides of
the interface, and more energy is reflected the greater the contrast.

The reflection coefficient (R) is a numerical measure of the effect of an interface on wave
propagation, and is calculated as the ratio of the amplitude A1 of the reflected ray to the
amplitude A0 of the incident ray.
A1
R=
A0

To relate this simple measure to the physical properties of the materials at the interface is a
complex problem. As we have already seen, the propagation of a P-wave depends on the bulk
and shear elastic moduli, as well as the density of the material. At the boundary the stress and
strain in the two materials must be considered. Since the materials are different, the relations
between stress and strain in each will be different. The orientation of stress and strain to the
interface also becomes important. The formal solution of this physical problem was derived early
in the 20th century, and the resulting equations are named the Zoeppritz equations (Zoeppritz
1919; and for explanation of derivations see Sheriff & Geldart 1982). Here, the solutions of these
equations will be accepted. For a normally incident ray the relationships are fairly simple, giving:
ρ2 v 2 −ρ1 v 1 Z 2−Z 1
R= =
ρ2 v 2 + ρ1 v 1 Z 2+ Z 1

Where ρ1, v1, Z1 and ρ2, v2, Z2 are the density, P-wave velocity and acoustic impedance values in
the first and second layers, respectively.
Therefore, the magnitude and polarity of the reflection coefficient depends on the difference
between seismic impedances of the two layers (Z 1 and Z2). Large differences (Z2– Z1) in seismic
impedances results in relatively large reflection coefficients. From the above equation the value
of R lies between -1 and 1.

i) Reflection and refraction of obliquely incident rays

Oblique incident rays are rays strikes an interface at an angle other than 90 0. When a P-wave ray
is obliquely incident on an interface of acoustic impedance contrast, reflected and transmitted P-
wave rays are generated as in the case of normal incidence. Additionally, some of the incident
compressional energy is converted into reflected and transmitted S-wave rays (Fig. 3.14) that are
polarized in a vertical plane. The Zoeppritz’ equations are a relatively complex set of equations
that allow calculation of the amplitudes of the two reflected and the two transmitted waves as
functions of the angle of incidence. The equations require P- and S-wave velocities (VP2, VS2,
VP1, and VS1 in Fig. 3.14) plus densities on both sides of the boundary. The S-waves that are
called converted rays contain information that can help identify fractured zones in reservoir
rocks. But here consideration will be confined to the P-waves.

Figure-3.14. Reflected and refracted P- and S-wave rays generated by a P-wave ray obliquely
incident on an interface of acoustic impedance contrast.

Snell’s Law

This relationship was originally developed in the study of optics. It does, however, apply equally
well to seismic waves. Its major application is to determine angles of reflection and refraction
from the incidence of seismic waves on layer boundaries and velocities of the layers.
Snell defined the ray parameter p = sin i/v, where i is the angle of inclination of the ray in a layer
in which it is travelling with a velocity v. The generalized form of Snell’s Law states that, along
any one ray, the ray parameter remains a constant. For the refracted P-wave ray shown in Fig.
3.15, therefore

sin θ1 sin θ 2 sin θ1 v1


= =
v1 v2 or
sin θ2 v2

Figure- 3.15 Reflected and refracted P-wave rays associated with a P-wave ray obliquely
incident on an interface of acoustic impedance contrast.

Snell’s law of reflection states that the angle at which a ray is reflected is equal to the angle of
incidence.

The portion of incident energy that is transmitted through the boundary and into the second layer
with changed direction of propagation is called a refracted ray. The direction of the refracted ray
depends upon the ratio of the velocities in the two layers. If an incidence ray goes from a
material with high velocity to a material with low velocity, then the refracted wave bends toward
the vertical. Conversely, if an incidence ray goes from a material with low velocity to a material
with high velocity, the refracted wave bend toward the horizontal.
Figure-3.16 shows propagation of seismic waves in layered medium a) from high velocity to low
velocity layer b) from low velocity to high velocity layer.

Imaging Earth’s Interior

Seismology is the study of vibrations within Earth. These vibrations are caused by events such as
earthquakes, extraterrestrial impacts, explosions, storm waves hitting the shore, and tides.
Seismology is applied to the detection and study of earthquakes, but seismic waves also provide
important information about Earth’s interior.

Seismic waves travel through different materials at different speeds, and we can apply
knowledge of how they interact with different materials to understand Earth’s layers and internal
structures. Similar to the way that ultrasound is used to image the human body, we can measure
how long it takes for seismic waves to travel from their source to a recording station.

Another feature of seismic waves is that some, called P-waves, can travel rapidly though both
liquids and solids, but others, called S-waves, can only travel though solids, and are slower than
P-waves. Observing where P-waves travel, and S-waves do not, allows us to identify regions
within Earth that are melted.

Seismic Wave Paths


Seismic waves travel in all directions from their source, but it is convenient to imagine the path
traced by one point on the wave front, and represent that path as a seismic ray (arrows, Figure
3.6).
Figure 3.6 Seismic waves and seismic rays. The paths of seismic waves can be represented as
rays. Seismic ray paths are bent when they enter a rock layer with a different seismic velocity.

When seismic waves encounter a different rock layer, some might bounce off the layer,
or reflect, as in the bottom layer of Figure 3.6. But some waves will travel through the layer. If
the wave travels at a different speed in the new layer, its path will be bent, or refracted, as it
crosses into the new layer. If the wave can travel faster in the new layer, it will be bent slightly
toward the contact between the two layers. In Figure 3.6, the ray can travel progressively faster
in each layer as it goes down through the layers, and it is bent slightly upward each time it
crosses into the next layer. The reverse happens if the wave slows down. On the right side of the
diagram, the wave is moving upward through slower and slower layers. It is bent away from the
faster layer each time, causing it to take a more direct path to the surface.

Seismic velocities are higher in more rigid layers, so broadly speaking, they get faster deeper
within Earth, because higher pressures make layers more rigid. They tend to take curved paths
through the Earth because refraction bends their path until they are reflected and directed upward
again, as in Figure 3.6.

Discoveries with Seismic Waves

The Moho: Where Crust Meets Mantle


One of the first discoveries about Earth’s interior made through seismology was in the early
1900s by Croatian seismologist Andrija Mohorovičić (pronounced Moho-ro-vi-chich). He
noticed that sometimes, seismic waves arrived at seismic stations (measuring locations) farther
from an earthquake before they arrived at closer ones. He reasoned that the waves that traveled
farther were faster because they bent down and traveled faster through different rocks (those of
the mantle) before being bent upward back into the crust (Figure 3.7).

Figure 3.7 Depiction of seismic waves emanating from an earthquake (red star). Some waves
travel through the crust to the seismic station (at ~6 km/s), while others go down into the mantle
(where they travel at ~8 km/s) and are bent upward toward the surface, reaching the station
before the ones that travelled only through the crust. Source: Steven Earle (2016) CC BY
4.0 view source

The boundary between the crust and the mantle is now known as the Mohorovičić
discontinuity (or Moho). Its depth is between 60 – 80 km beneath major mountain ranges, 30 –
50 km beneath most of the continental crust, and 5 – 10 km beneath ocean crust.

The Core-Mantle Boundary


Arguments for a liquid outer core were supported by a distinctive signature in the global
distribution of seismic waves from earthquakes. When an earthquake occurs, there is a zone on
the opposite side of Earth where S-waves are not measured. This S-wave shadow zone begins
103° on either side of the earthquake, for a total angular distance of 154° (Figure 3.8, left). There
is also a P-wave shadow zone on either side of the earthquake, from 103° to 150° (Figure 3.8,
right).
Figure 3.8 Patterns of seismic wave propagation through Earth’s mantle and core. S-waves do
not travel through the liquid outer core, so they leave a shadow on Earth’s far side. P-waves do
travel through the core, but because the waves that enter the core are refracted, there are also P-
wave shadow zones. Source: Steven Earle (2016) CC BY 4.0 view source

The S-wave shadow zone occurs because S-waves cannot travel through the liquid outer core.
The P-wave shadow zone occurs because seismic velocities are much lower in the liquid outer
core than in the overlying mantle, and the P-waves are refracted in a way that leaves a gap. Not
only do the shadow zones tell us that the outer core is liquid, the size of the shadow zones allows
us to calculate the size of the core, and the location of the core-mantle boundary.

Seismic waves in the Earth’s Layers


The change seismic wave velocity with depth in Earth (Figure 3.9) has been determined over the
past several decades by analyzing seismic signals from large earthquakes all around the world.
Earth’s layers are detectable as changes in velocity with depth. The asthenosphere is visible as a
low velocity zone within the upper mantle (Figure 3.9, left). There is an abrupt increase in P-
wave velocity at 420 km, showing the depth at which minerals transform into structures that are
more stable at higher pressures and temperatures.
Figure 3.9 P-wave and S-wave velocity variations with depth from the crust through the upper
mantle (left) and from the crust through to the core (right). Source: Karla Panchuk (2018) CC
BY 4.0, modified after Steven Earle (2016) CC BY 4.0 view source left/ right

The boundary between the upper and lower mantle is visible at 660 km as a sudden change from
rapidly increasing P- and S-wave velocities to slow or no change in P-wave and S-wave
velocities (Figure 3.9, right). The core-mantle boundary (CMB in Figure 3.9) is apparent as a
sudden drop in P-wave velocities, where seismic waves move from solid mantle to liquid outer
core. The boundary between the outer core and inner core is marked by a sudden increase in P-
wave velocity after 5000 km, where seismic waves move from a liquid back into a solid again.

Seismic Images of Plate Tectonic Structures


Using data from many seismometers and hundreds of earthquakes, it is possible to create images
from the seismic properties of the mantle. This technique is known as seismic tomography.
Tomography can be used to map out slabs of lithosphere that are entering the mantle, or have
disappeared within it. Those slabs are cooler, and therefore more rigid than surrounding mantle
rocks, so seismic waves travel through them faster. In Figure 3.10, higher-than-average seismic
velocities in cool slabs are indicated in dark blue.

Figure 3.10 P-waves and S-waves used to map out the location of the Cocos slab of lithosphere.
The slab appears in dark blue, indicating higher than average seismic wave velocities. Left-
Tomograms showing seismic wave anomalies for a 1290 km surface. Right- Cross-sections
along the transect marked X-Y on the globe. Source: Karla Panchuk (2018) CC BY 4.0, modified
after van der Meer et al. (2018) CC BY 4.0 view source

Thanks to the tomograms, we can see that the Cocos plate, which is colliding with Central
America, is part of a much larger slab of lithosphere that has already settled onto the mantle.
Tomograms representing a surface at 1290 km depth (Figure 3.10, left) show that at that level,
the Cocos slab is beneath the Caribbean Sea. The tomograms on the right show a vertical view
along the line X-Y marked on the globe. The vertical tomograms show us that the Cocos slab
extends all the way down to the core-mantle boundary.
Earth’s Interior Heat

Earth Gets Hotter the Deeper


Earth’s temperature increases with depth, but not at a uniform rate (Figure 3.11).
Earth’s geothermal gradient is 15° to 30°C/km within the crust. It then drops off dramatically
through the mantle, increases more quickly at the base of the mantle, and then increases slowly
through the core. The temperature is approximately 1000°C at the base of the crust, around
3500°C at the base of the mantle, and approximately 6,000°C at Earth’s centre.

Figure 3.11 Geothermal gradient (change in temperature with depth). Left- Geothermal gradient
in the crust and upper mantle. The geothermal gradient remains below the melting temperature of
rock, except in the asthenosphere. There, temperatures are high enough to melt some of the
minerals. Right- Geothermal gradient throughout Earth. Rapid changes occur in the uppermost
mantle, and at the core-mantle boundary. Source: Karla Panchuk (2018) CC BY 4.0, modified
after Steven Earle (2016) CC BY 4.0 view source left/ right
The temperature gradient within the lithosphere varies depending on the tectonic setting.
Gradients are lowest in the central parts of continents, higher where plates collide, and higher
still at boundaries where plates are moving away from each other.

In spite of high temperatures within Earth, mantle rocks are almost entirely solid. High pressures
keep them from melting. The red dashed line in Figure 3.11 (right) shows the minimum
temperature at which dry mantle rocks will melt. Rocks at temperatures to the left of the line will
remain solid. In rocks at temperatures to the right of the line, some minerals will begin to melt.
Notice that the red dashed line goes further to the right for greater depths, and therefore greater
pressures. Now compare the geothermal gradient with the red dashed line. The geothermal
gradient is to the left of the red line, except in the asthenosphere, where small amounts of melt
are present.

Convection Helps to Move Heat Within Earth


The fact that the temperature gradient is much lower in the main part of the mantle than in the
lithosphere has been interpreted as evidence of convection in the mantle. When the mantle
convects, heat is transferred through the mantle by physically moving hot rocks. Mantle
convection is the result of heat transfer from the core to the base of the lower mantle. As with a
pot of soup on a hot stove (Figure 3.12), the material near the heat source (the soup at the bottom
of the pot) becomes hot and expands, making it less dense than the material above. Buoyancy
causes it to rise, and cooler material flows in from the sides. Of course, convection in the soup
pot is much faster than convection in the mantle. Mantle convection occurs at rates of
centimetres per year.

Figure 3.12 Convection in a pot of soup on a hot stove (left). As long as heat is being transferred
from below, the liquid will convect. If the heat is turned off (right), the liquid remains hot for a
while, but convection will cease. Source: Steven Earle (2015) CC-BY 4.0 view source

Convection carries heat to the surface of the mantle much faster than heating by conduction.
Conduction is heat transfer by collisions between molecules, and is how heat is transferred from
the stove to the soup pot. A convecting mantle is an essential feature of plate tectonics, because
the higher rate of heat transfer is necessary to keep the asthenosphere weak. Earth’s mantle will
stop convecting once the core has cooled to the point where there is not enough heat transfer to
overcome the strength of the rock. This has already happened on smaller planets like Mercury
and Mars, as well as on Earth’s moon. When mantle convection stops, the end of plate tectonics
will follow.

Models of Mantle Convection


In the soup pot example, convection moves hot soup from the bottom of the pot to the top. Some
geologists think that Earth’s convection works the same way— hot rock from the base of the
mantle moves all the way to the top of the mantle before cooling and sinking back down again.
This view is referred to as whole-mantle convection (Figure 3.8, left). Other geologists think
that the upper and lower mantle are too different to convect as one. They point to slabs of
lithosphere that are sinking back into the mantle, some of which seem to perch on the boundary
between the upper and lower mantle, rather than sinking straight through. They also note
chemical differences in magma originating in different parts of the mantle— differences that are
not consistent with the entire mantle being well stirred. They argue that double-layered
convection is a better fit with the observations (Figure 3.13, right). Still others argue that there
may be some locations where convection goes from the bottom of the mantle to the top, and
some where it doesn’t (Figure 3.13, middle).

Figure 3.13 Models of mantle convection. Left- whole mantle convection. Rocks rise from the
core-mantle boundary to the top of the mantle, then sink to the bottom again. Right- Two-layer
convection, in which upper and lower mantle convect at different rates. Middle- Convection
paths vary depending on the circumstances. Source: Karla Panchuk (2018) CC BY 4.0

Why Is Earth Hot Inside?


The heat of Earth’s interior comes from a variety of sources. These include the heat contained in
the objects that accreted to form Earth, and the heat produced when they collided. As Earth grew
larger, the increased pressure on Earth’s interior caused it to compress and heat up. Heat also
came from friction when melted material was redistributed within Earth, forming the core and
mantle.

A major source of Earth’s heat is radioactivity, the energy released when the unstable atoms
decay. The radioactive isotopes uranium-235 ( 235U), uranium-238 (238U), potassium-40 (40K), and
thorium-232 (232Th) in Earth’s mantle are the primary source. Radioactive decay produced more
heat early in Earth’s history than it does today, because fewer atoms of those isotopes are left
today (Figure 3.14). Heat contributed by radioactivity is now roughly a quarter what it was when
Earth formed.

Figure 3.14 Production of heat within the Earth over time by radioactive decay of uranium,
thorium, and potassium. Heat production has decreased over time as the abundance of
radioactive atoms has decreased. Source: Steven Earle (2015) CC BY 4.0 view source, modified
after Arevalo et al. (2009)

References
Arevalo, R., McDonough, W., & Luong, M. (2009). The K/U ratio of Earth: Insights into mantle
composition, structure and thermal evolution. Earth and Planetary Science Letters, 278(3-4),
361-369. https://doi.org/10.1016/j.epsl.2008.12.023

Exercise 9.1 How soon will seismic waves get here?


Imagine that a strong earthquake takes place on Vancouver Island within Strathcona Park
(west of Courtenay). Assuming that the crustal average P wave velocity is 5 km per second,
how long will it take (in seconds) for the first seismic waves (P waves) to reach you in the
following places (distances from the epicentre are shown)?

1. Nanaimo (120 km away)


2. Surrey (200 km away)
3. Kamloops (390 km away)

See Appendix 3 for Exercise 9.1 answers.


Times shown for velocity of 5 kilometres per second (km/s).

1. Nanaimo (120 kilometres), 24 seconds.


2. Surrey (200 kilometres), 40 seconds.
3. Kamloops (390 kilometres), 78 seconds

You might also like