You are on page 1of 21

CE 232 Earthquake Engineering

SEISMOLOGY AND EARTHQUAKE REPORT

Submitted by

Eduardo M. Oroz Jr.

TABLE OF CONTENTS

Internal Structure of the Earth.3


Continental Drift and Plate Tectonics..5
Rupture of fault and Earthquakes.7
Other sources of Earthquakes..10
Location of Earthquakes.12
Size and Measurement of Earthquakes13
Modification of Earthquake due to the Nature of Soil..16
Earthquake Damage to Structures..17

Internal Structure of the Earth


Three Parts of Earth's Interior
Knowledge of earth's interior is essential for understanding plate tectonics. A good
analogy for teaching about earth's interior is a piece of fruit with a large pit such as a peach or a
plum. Most students are familiar with these fruits and have seen them cut in half.
If we cut a piece of fruit in half we will see that it is composed of three parts: 1) a very
thin skin, 2) a seed of significant size located in the center, and 3) most of the mass of the fruit
being contained within the flesh. Cutting the earth we would see: 1) a very thin crust on the
outside, 2) a core of significant size in the center, and 3) most of the mass of the Earth contained
in the mantle.

Earth's Crust
There are two different types of crust: thin oceanic crust that underlies the ocean basins
and thicker continental crust that underlies the continents. These two different types of crust are
made up of different types of rock. The thin oceanic crust is composed of primarily of basalt and
the thicker continental crust is composed primarily of granite. The low density of the thick
continental crust allows it to "float" in high relief on the much higher density mantle below.

Earth's Mantle
Earth's mantle is thought to be composed mainly of olivine-rich rock. It has different
temperatures at different depths. The temperature is lowest immediately beneath the crust and
increases with depth. The highest temperatures occur where the mantle material is in contact
with the heat-producing core. This steady increase of temperature with depth is known as the
geothermal gradient. The geothermal gradient is responsible for different rock behaviors and the
different rock behaviors are used to divide the mantle into two different zones. Rocks in the
upper mantle are cool and brittle, while rocks in the lower mantle are hot and soft (but not
molten). Rocks in the upper mantle are brittle enough to break under stress and produce
earthquakes. However, rocks in the lower mantle are soft and flow when subjected to forces
instead of breaking. The lower limit of brittle behavior is the boundary between the upper and
lower mantle.

Earth's Core
Earth's Core is thought to be composed mainly of an iron and nickel alloy. This
composition is assumed based upon calculations of its density and upon the fact that many
meteorites are iron-nickel alloys. The core is earth's source of internal heat because it contains
radioactive materials which release heat as they break down into more stable substances.
The core is divided into two different zones. The outer core is a liquid because the
temperatures there are adequate to melt the iron-nickel alloy. However, the inner core is a solid
even though its temperature is higher than the outer core. Here, tremendous pressure, produced
by the weight of the overlying rocks is strong enough to crowd the atoms tightly together and
prevents the liquid state.
4

Continental Drift and Plate Tectonics


Continental Drift
Continental drift is the movement of the Earth's continents relative to each other, thus
appearing to drift across the ocean bed. The speculation that continents might have 'drifted' was
first put forward by Abraham Ortelius in 1596. The concept was independently and more fully
developed by Alfred Wegener in 1912, but his theory was rejected by some for lack of a
mechanism (though this was supplied later by Holmes) and others because of prior theoretical
commitments. The idea of continental drift has been subsumed by the theory of plate tectonics,
which explains how the continents move.
Plate Tectonics
The continents drift slowly (the timescale for substantial change is 10-100 million years),
but that they drift at all is remarkable. The following figure illustrates the structure of the first
100-200 kilometers of the Earth's interior, and provides an answer to this question.
The crust is thin, varying from a few tens of kilometers thick beneath the continents to to
less than 10 km thick beneath the many of the oceans. The crust and upper mantle together
constitute the lithosphere, which is typically 50-100 km thick and is broken into large plates (not
illustrated). These plates sit on the aesthenosphere.
The aesthenosphere is kept plastic (deformable) largely through heat generated by
radioactive decay. The material that is decaying is primarily radioactive isotopes of light
elements like aluminum and magnesium. This heat source is small on an absolute scale (the
corresponding heat flow at the surface out of the Earth is only about 1/6000 of the solar energy

falling on the surface). Nevertheless, because of the insulating properties of the Earth's rocks this
is sufficient to keep the aesthenosphere plastic in consistency.
Plate tectonics (from the Late Latin tectonicus, from the Greek: "pertaining to
building") is a scientific theory that describes the large-scale motion of Earth's lithosphere. This
theoretical model builds on the concept of continental drift which was developed during the first
few decades of the 20th century. The geoscientific community accepted the theory after the
concepts of seafloor spreading were later developed in the late 1950s and early 1960s.

The lithosphere and the aesthenosphere

The lithosphere, which is the rigid outermost shell of a planet (on Earth, the crust and
upper mantle), is broken up into tectonic plates. On Earth, there are seven or eight major plates
(depending on how they are defined) and many minor plates. Where plates meet, their relative
motion determines the type of boundary; convergent, divergent, or transform. Earthquakes,
volcanic activity, mountain-building, and oceanic trench formation occur along these plate
boundaries. The lateral relative movement of the plates typically varies from zero to 100 mm
annually.
Tectonic plates are composed of oceanic lithosphere and thicker continental lithosphere,
each topped by its own kind of crust. Along convergent boundaries, subduction carries plates
into the mantle; the material lost is roughly balanced by the formation of new (oceanic) crust
along divergent margins by seafloor spreading. In this way, the total surface of the globe remains
the same. This prediction of plate tectonics is also referred to as the conveyor belt principle.

Earlier theories (that still have some supporters) propose gradual shrinking (contraction) or
gradual expansion of the globe.
Tectonic plates are able to move because the Earth's lithosphere has greater strength than
the underlying asthenosphere. Lateral density variations in the mantle result in convection. Plate
movement is thought to be driven by a combination of the motion of the seafloor away from the
spreading ridge (due to variations in topography and density of the crust, which result in
differences in gravitational forces) and drag, with downward suction, at the subduction zones.
Another explanation lies in the different forces generated by the rotation of the globe and the
tidal forces of the Sun and Moon. The relative importance of each of these factors and their
relationship to each other is unclear, and still the subject of much debate.
Rupture of Fault and Earthquakes
Rupture of Faults
A fault is a break in the earth's crust along which movement can take place causing an
earthquake. In Utah, movement along faults is mostly vertical; mountain blocks (for example, the
Wasatch Range) move up relative to the downward movement of valley blocks (for example, the
Salt Lake Valley).
Faults with evidence of Holocene (about 10,000 years ago to present) movement are the
main concern because they are most likely to generate future earthquakes. If the earthquake is
large enough, surface fault rupture can occur.
With a large earthquake (about magnitude 6.5 and greater), the fault rupture can reach
and displace the ground surface, forming a fault scarp (steep break in slope). The resulting fault
scarp may be several inches to 20 feet in height, and up to about 40 miles in length, depending
on the size of the earthquake.
An area hundreds of feet wide can be affected, called the zone of deformation, which
occurs chiefly on the downthrown side of the main fault and encompasses multiple minor faults,
cracks, local tilting, and grabens (downdropped blocks between faults). Buildings in the zone of
deformation would be damaged, particularly those straddling the main fault.
Also, anything crossing the fault, such as transportation corridors, utilities, and other
lifelines, both underground and above ground, can be damaged or broken. The ground can be
dropped below the water table on the downdropped side, resulting in localized flooding.
Surface fault rupture can also cause tectonic subsidence, which is the broad, permanent
tilting of the valley floor down toward the fault scarp. Tilting can cause flooding along lake and
reservoir shorelines nearest the fault; along altered stream courses; and along canals, sewer lines,
or other gravity-flow systems where slope gradients are lessened or reversed.

Earthquake fault types


There are three main types of fault, all of which may cause an interplate earthquake:
normal, reverse (thrust) and strike-slip. Normal and reverse faulting are examples of dip-slip,
where the displacement along the fault is in the direction of dip and movement on them involves
a vertical component. Normal faults occur mainly in areas where the crust is being extended such
as a divergent boundary. Reverse faults occur in areas where the crust is being shortened such as
at a convergent boundary. Strike-slip faults are steep structures where the two sides of the fault
slip horizontally past each other; transform boundaries are a particular type of strike-slip fault.
Many earthquakes are caused by movement on faults that have components of both dip-slip and
strike-slip; this is known as oblique slip.
Reverse faults, particularly those along convergent plate boundaries are associated with
the most powerful earthquakes, megathrust earthquakes, including almost all of those of
magnitude 8 or more. Strike-slip faults, particularly continental transforms can produce major
earthquakes up to about magnitude 8. Earthquakes associated with normal faults are generally
less than magnitude 7.
This is so because the energy released in an earthquake, and thus its magnitude, is
proportional to the area of the fault that ruptures and the stress drop. Therefore, the longer the
length and the wider the width of the faulted area, the larger the resulting magnitude. The
topmost, brittle part of the Earth's crust, and the cool slabs of the tectonic plates that are
descending down into the hot mantle, are the only parts of our planet which can store elastic
energy and release it in fault ruptures. Rocks hotter than about 300 degrees Celsius flow in
response to stress; they do not rupture in earthquakes. The maximum observed lengths of
ruptures and mapped faults, which may break in one go are approximately 1000 km. Examples
are the earthquakes in Chile, 1960; Alaska, 1957; Sumatra, 2004, all in subduction zones. The
longest earthquake ruptures on strike-slip faults, like the San Andreas Fault (1857, 1906), the
North Anatolian Fault in Turkey (1939) and the Denali Fault in Alaska (2002), are about half to
one third as long as the lengths along subducting plate margins, and those along normal faults are
even shorter.
The most important parameter controlling the maximum earthquake magnitude on a fault
is however not the maximum available length, but the available width because the latter varies by
a factor of 20. Along converging plate margins, the dip angle of the rupture plane is very
shallow, typically about 10 degrees. Thus the width of the plane within the top brittle crust of the
Earth can become 50 to 100 km (Japan, 2011; Alaska, 1964), making the most powerful
earthquakes possible.
Strike-slip faults tend to be oriented near vertically, resulting in an approximate width of
10 km within the brittle crust, thus earthquakes with magnitudes much larger than 8 are not
possible. Maximum magnitudes along many normal faults are even more limited because many
8

of them are located along spreading centers, as in Iceland, where the thickness of the brittle layer
is only about 6 km.
In addition, there exists a hierarchy of stress level in the three fault types. Thrust faults
are generated by the highest, strike slip by intermediate and normal faults by the lowest stress
levels. This can easily be understood by considering the direction of the greatest principal stress,
the direction of the force that 'pushes' the rock mass during the faulting. In the case of normal
faults, the rock mass is pushed down in a vertical direction, thus the pushing force (greatest
principal stress) equals the weight of the rock mass itself. In the case of thrusting, the rock mass
'escapes' in the direction of the least principal stress, namely upward, lifting the rock mass up,
thus the overburden equals the least principal stress. Strike-slip faulting is intermediate between
the other two types described above. This difference in stress regime in the three faulting
environments can contribute to differences in stress drop during faulting, which contributes to
differences in the radiated energy, regardless of fault dimensions.
Earthquake
An earthquake (also known as a quake, tremor or temblor) is the result of a sudden
release of energy in the Earth's crust that creates seismic waves. The seismicity, seismism or
seismic activity of an area refers to the frequency, type and size of earthquakes experienced over
a period of time.
Earthquakes are measured using observations from seismometers. The moment
magnitude is the most common scale on which earthquakes larger than approximately 5 are
reported for the entire globe. The more numerous earthquakes smaller than magnitude 5 reported
by national seismological observatories are measured mostly on the local magnitude scale, also
referred to as the Richter magnitude scale. These two scales are numerically similar over their
range of validity. Magnitude 3 or lower earthquakes are mostly almost imperceptible or weak
and magnitudes 7 and over potentially cause serious damage over larger areas, depending on
their depth. The largest earthquakes in historic times have been of magnitude slightly over 9,
although there is no limit to the possible magnitude. The most recent large earthquake of
magnitude 9.0 or larger was a 9.0 magnitude earthquake in Japan in 2011 (as of March 2014),
and it was the largest Japanese earthquake since records began. Intensity of shaking is measured
on the modified Mercalli scale. The shallower an earthquake, the more damage to structures it
causes, all else being equal.
At the Earth's surface, earthquakes manifest themselves by shaking and sometimes
displacement of the ground. When the epicenter of a large earthquake is located offshore, the
seabed may be displaced sufficiently to cause a tsunami. Earthquakes can also trigger landslides,
and occasionally volcanic activity.
In its most general sense, the word earthquake is used to describe any seismic event
whether natural or caused by humans that generates seismic waves. Earthquakes are caused
9

mostly by rupture of geological faults, but also by other events such as volcanic activity,
landslides, mine blasts, and nuclear tests. An earthquake's point of initial rupture is called its
focus or hypocenter. The epicenter is the point at ground level directly above the hypocenter.
Other Sources of Earthquakes
Naturally occurring earthquakes
Tectonic earthquakes occur anywhere in the earth where there is sufficient stored elastic
strain energy to drive fracture propagation along a fault plane. The sides of a fault move past
each other smoothly and aseismically only if there are no irregularities or asperities along the
fault surface that increase the frictional resistance. Most fault surfaces do have such asperities
and this leads to a form of stick-slip behaviour. Once the fault has locked, continued relative
motion between the plates leads to increasing stress and therefore, stored strain energy in the
volume around the fault surface. This continues until the stress has risen sufficiently to break
through the asperity, suddenly allowing sliding over the locked portion of the fault, releasing the
stored energy. This energy is released as a combination of radiated elastic strain seismic waves,
frictional heating of the fault surface, and cracking of the rock, thus causing an earthquake. This
process of gradual build-up of strain and stress punctuated by occasional sudden earthquake
failure is referred to as the elastic-rebound theory. It is estimated that only 10 percent or less of
an earthquake's total energy is radiated as seismic energy. Most of the earthquake's energy is
used to power the earthquake fracture growth or is converted into heat generated by friction.
Therefore, earthquakes lower the Earth's available elastic potential energy and raise its
temperature, though these changes are negligible compared to the conductive and convective
flow of heat out from the Earth's deep interior.
Earthquakes away from plate boundaries
Where plate boundaries occur within the continental lithosphere, deformation is spread
out over a much larger area than the plate boundary itself. In the case of the San Andreas fault
continental transform, many earthquakes occur away from the plate boundary and are related to
strains developed within the broader zone of deformation caused by major irregularities in the
fault trace (e.g., the "Big bend" region). The Northridge earthquake was associated with
movement on a blind thrust within such a zone. Another example is the strongly oblique
convergent plate boundary between the Arabian and Eurasian plates where it runs through the
northwestern part of the Zagros mountains. The deformation associated with this plate boundary
is partitioned into nearly pure thrust sense movements perpendicular to the boundary over a wide
zone to the southwest and nearly pure strike-slip motion along the Main Recent Fault close to the
actual plate boundary itself. This is demonstrated by earthquake focal mechanisms.
All tectonic plates have internal stress fields caused by their interactions with
neighbouring plates and sedimentary loading or unloading (e.g. deglaciation). These stresses may
be sufficient to cause failure along existing fault planes, giving rise to intraplate earthquakes.
10

Shallow-focus and deep-focus earthquakes


The majority of tectonic earthquakes originate at the ring of fire in depths not exceeding
tens of kilometers. Earthquakes occurring at a depth of less than 70 km are classified as 'shallowfocus' earthquakes, while those with a focal-depth between 70 and 300 km are commonly termed
'mid-focus' or 'intermediate-depth' earthquakes. In subduction zones, where older and colder
oceanic crust descends beneath another tectonic plate, deep-focus earthquakes may occur at
much greater depths (ranging from 300 up to 700 kilometers). These seismically active areas of
subduction are known as Wadati-Benioff zones. Deep-focus earthquakes occur at a depth where
the subducted lithosphere should no longer be brittle, due to the high temperature and pressure.
A possible mechanism for the generation of deep-focus earthquakes is faulting caused by olivine
undergoing a phase transition into a spinel structure.
Rupture dynamics
A tectonic earthquake begins by an initial rupture at a point on the fault surface, a process
known as nucleation. The scale of the nucleation zone is uncertain, with some evidence, such as
the rupture dimensions of the smallest earthquakes, suggesting that it is smaller than 100 m while
other evidence, such as a slow component revealed by low-frequency spectra of some
earthquakes, suggest that it is larger. The possibility that the nucleation involves some sort of
preparation process is supported by the observation that about 40% of earthquakes are preceded
by foreshocks. Once the rupture has initiated it begins to propagate along the fault surface. The
mechanics of this process are poorly understood, partly because it is difficult to recreate the high
sliding velocities in a laboratory. Also the effects of strong ground motion make it very difficult
to record information close to a nucleation zone.
Rupture propagation is generally modeled using a fracture mechanics approach, likening
the rupture to a propagating mixed mode shear crack. The rupture velocity is a function of the
fracture energy in the volume around the crack tip, increasing with decreasing fracture energy.
The velocity of rupture propagation is orders of magnitude faster than the displacement velocity
across the fault. Earthquake ruptures typically propagate at velocities that are in the range 70
90% of the S-wave velocity and this is independent of earthquake size. A small subset of
earthquake ruptures appear to have propagated at speeds greater than the S-wave velocity. These
supershear earthquakes have all been observed during large strike-slip events. The unusually
wide zone of coseismic damage caused by the 2001 Kunlun earthquake has been attributed to the
effects of the sonic boom developed in such earthquakes. Some earthquake ruptures travel at
unusually low velocities and are referred to as slow earthquakes. A particularly dangerous form
of slow earthquake is the tsunami earthquake, observed where the relatively low felt intensities,
caused by the slow propagation speed of some great earthquakes, fail to alert the population of
the neighbouring coast, as in the 1896 Meiji-Sanriku earthquake.

11

Tidal forces
Research work has shown a robust correlation between small tidally induced forces and nonvolcanic tremor activity.
Earthquake clusters
Most earthquakes form part of a sequence, related to each other in terms of location and
time. Most earthquake clusters consist of small tremors that cause little to no damage, but there is
a theory that earthquakes can recur in a regular pattern.
Aftershocks
An aftershock is an earthquake that occurs after a previous earthquake, the mainshock.
An aftershock is in the same region of the main shock but always of a smaller magnitude. If an
aftershock is larger than the main shock, the aftershock is redesignated as the main shock and the
original main shock is redesignated as a foreshock. Aftershocks are formed as the crust around
the displaced fault plane adjusts to the effects of the main shock.
Earthquake swarms
Earthquake swarms are sequences of earthquakes striking in a specific area within a short
period of time. They are different from earthquakes followed by a series of aftershocks by the
fact that no single earthquake in the sequence is obviously the main shock, therefore none have
notable higher magnitudes than the other. An example of an earthquake swarm is the 2004
activity at Yellowstone National Park. In August 2012, a swarm of earthquakes shook Southern
California's Imperial Valley, showing the most recorded activity in the area since the 1970s.
Earthquake storms
Sometimes a series of earthquakes occur in a sort of earthquake storm, where the
earthquakes strike a fault in clusters, each triggered by the shaking or stress redistribution of the
previous earthquakes. Similar to aftershocks but on adjacent segments of fault, these storms
occur over the course of years, and with some of the later earthquakes as damaging as the early
ones. Such a pattern was observed in the sequence of about a dozen earthquakes that struck the
North Anatolian Fault in Turkey in the 20th century and has been inferred for older anomalous
clusters of large earthquakes in the Middle East.
Location of Earthquakes
The primary purpose of a seismometer is to locate the initiating points of earthquake
epicenters. The secondary purpose, of determining the 'size' or Moment magnitude scale must be
calculated after the precise location is known.

12

The earliest seismographs were designed to give a sense of the direction of the first
motions from an earthquake. The Chinese frog seismograph would have dropped its ball in the
general compass direction of the earthquake, assuming a strong positive pulse. We now know
that first motions can be in almost any direction depending on the type of initiating rupture (focal
mechanism).
The first refinement that allowed a more precise determination of the location, was the
use of a time scale. Instead of merely noting, or recording, the absolute motions of a pendulum,
the displacements were plotted on a moving graph, driven by a clock mechanism. This was the
first seismogram, which allowed precise timing of the first ground motion, and an accurate plot
of subsequent motions.
From the first seismograms, as seen on the figure, it was noticed that the trace was
divided into two major portions. The first seismic wave to arrive was the P-wave, followed
closely by the S-wave. Knowing the relative 'velocities of propagation', it was a simple matter to
calculate the distance of the earthquake.
One seismograph would give the distance, but that could be plotted as a circle, with an
infinite number of possibilities. Two seismographs would give two intersecting circles, with two
possible locations. Only with a third seismograph would there be a precise location.
The process of accurate location, was greatly improved with the advent of precise
absolute timing. Early seismographs were almost always located at an astronomical observatory,
just for the purpose of timing. See the history of the Canadian Dominion Observatory, is also the
Geological Survey of Canada seismology laboratory. Recently, GPS is being used for accurate
time, and seismometers can be located almost anywhere.
Modern earthquake location still requires a minimum of three seismometers. Most likely,
there are many, forming a seismic array. The emphasis is on precision, since much can be
learned about the fault mechanics and seismic hazard, if the locations can be determined to
within a kilometer or two, for small earthquakes. For this, computer programs use an iterative
process, involving a 'guess and correction' algorithm. As well, a very good model of the local
crustal velocity structure is required: seismic velocities vary with the local geology. For P-waves,
the relation between velocity and bulk density of the medium has been quantified in Gardner's
relation.
Size and Measurement of Earthquakes
Measuring and locating earthquakes
Earthquakes can be recorded by seismometers up to great distances, because seismic
waves travel through the whole Earth's interior. The absolute magnitude of a quake is
conventionally reported by numbers on the moment magnitude scale (formerly Richter scale,
13

magnitude 7 causing serious damage over large areas), whereas the felt magnitude is reported
using the modified Mercalli intensity scale (intensity IIXII).
Every tremor produces different types of seismic waves, which travel through rock with
different velocities:
Longitudinal P-waves (shock- or pressure waves)
Transverse S-waves (both body waves)
Surface waves (Rayleigh and Love waves)
Propagation velocity of the seismic waves ranges from approx. 3 km/s up to 13 km/s,
depending on the density and elasticity of the medium. In the Earth's interior the shock- or P
waves travel much faster than the S waves (approx. relation 1.7 : 1). The differences in travel
time from the epicentre to the observatory are a measure of the distance and can be used to image
both sources of quakes and structures within the Earth. Also the depth of the hypocenter can be
computed roughly.
In solid rock P-waves travel at about 6 to 7 km per second; the velocity increases within
the deep mantle to ~13 km/s. The velocity of S-waves ranges from 23 km/s in light sediments
and 45 km/s in the Earth's crust up to 7 km/s in the deep mantle. As a consequence, the first
waves of a distant earthquake arrive at an observatory via the Earth's mantle.
On average, the kilometer distance to the earthquake is the number of seconds between
the P and S wave times 8.[46] Slight deviations are caused by inhomogeneities of subsurface
structure. By such analyses of seismograms the Earth's core was located in 1913 by Beno
Gutenberg.
Earthquakes are not only categorized by their magnitude but also by the place where they
occur. The world is divided into 754 FlinnEngdahl regions (F-E regions), which are based on
political and geographical boundaries as well as seismic activity. More active zones are divided
into smaller F-E regions whereas less active zones belong to larger F-E regions.
Standard reporting of earthquakes includes its magnitude, date and time of occurrence,
geographic coordinates of its epicenter, depth of the epicenter, geographical region, distances to
population centers, location uncertainty, a number of parameters that are included in USGS
earthquake reports (number of stations reporting, number of observations, etc.), and a unique
event ID.
Size of Earthquakes
The magnitude is the most often cited measure of an earthquake's size, but it is not the
only measure, and in fact, there are different types of earthquake magnitude. Early estimates of
earthquake size were based on non-instrumental measures of the earthquakes effects. For
14

example, we could use values such as the number of fatalities or injuries, the maximum value of
shaking intensity, or the area of intense shaking. The problem with these measures is that they
don't correlate well. The damage and devastation produced by an earthquake will depend on its
location, depth, proximity to populated regions, as well as its "true" size. Even for earthquakes
close enough to population centers values such as maximum intensity and the area experiencing
a particular level of shaking did not correlate well.
In 1931 a Japanese seismologist named Kiyoo Wadati constructed a chart of maximum
ground motion versus distance for a number of earthquakes and noted that the plots for different
earthquakes formed parallel, curved lines (the larger earthquakes produced larger amplitudes).
The fact that earthquakes of different size generated curves that were roughly parallel suggested
that a single number could quantify the relative size of different earthquakes.
In 1935 Charles Richter constructed a similar diagram of peak ground motion versus
distance and used it to create the first earthquake magnitude scale (a logarithmic relationship
between earthquake size and observed peak ground motion). He based his scale on an analogy
with the stellar brightness scale commonly used in astronomy which is also similar to the pH
scale used to measure acidity (pH is a logarithmic measure of the Hydrogen ion concentration in
a solution).
To complete the construction of the magnitude scale, Richter had to establish a reference
value and identify the rate at which the peak amplitudes decrease with distance from an
earthquake. He established a reference value for earthquake magnitude when he defined the
magnitude as the base-ten logarithm of the maximum ground motion (in micrometers) recorded
on a Wood-Anderson short-period seismometer one hundred kilometers from the earthquake.
Richter was pragmatic in his definition, and chose a value for a magnitude zero that insured that
most of the earthquakes routinely recorded would have positive magnitudes. Also, the WoodAnderson short-period instrument that Richter chose for his reference records seismic waves
with a period of about 0.8 seconds, roughly the vibration periods that we feel and that damage
our buildings and other structures.
Richter also developed a distance correction to account for the variation in maximum
ground motion with distance from an earthquake (the dashed curves shown in the above diagram
show his relationship for southern California). The precise rate that the peak ground motions
decrease with distance depends on the regional geology and thus the magnitude scale for
different regions is slightly dependent on the "distance correction curve".
Thus originally, Richter's scale was specifically designed for application in southern
California. Richter's method became widely used because it was simple, required only the
location of the earthquake (to get the distance) and a quick measure of the peak ground motion,
was more reliable than older measures such as intensity. It became widely used, well established,
and forms the basis for many of the measures that we continue to use today. Generally the

15

magnitude is computed from seismographs from as many seismic recording stations as are
available and the average value is used as our estimate of an earthquake's size.
We call the Richter's original magnitude scale ML (for "local magnitude"), but the press
usually reports all magnitudes as Richter magnitudes.
Modification of Earthquake due to Nature of Soil
Soil Conditions can have considerable effect on the shape of the surface motion response
during earthquakes. Since the changes in the shape of the response spectra can have significant
effects on the lateral forces on structures, it would seem desirable that the nature of soil
conditions underlying a site be taken into account when evaluating the lateral forces for design
purposes. It is also apparent that the most accurate information available on the dynamic
properties of such soils should be used. While the information produced by Seed and Idriss has
been shown to differ from that determined for this study, the caution that their data is only
intended to be used as a guideline and considerable judgments may have to be exercised. They
suggest using a range of material properties in the computations when equipment for the accurate
determination of moduli and equivalent viscous damping factors is not available.
The various site properties are extremely involved in controlling the response, since with
increasing input acceleration the amplification factor decreases due to the higher damping
employed in the softer soils. The highest amplification factors are generally associated with the
weaker motions. This phenomenon has been observed by many researchers and writers.
One aspect which cannot be bypassed is the influence of the characteristics of the base
rock motion on ground response. Insufficient evidence is available to state categorically whether
bedrock motions should be characterized by response spectra having a flat or a peaked form.
Many writers believe the former to be a better representation. Regardless of the general form of
the spectra the fact should not be overlooked that seemingly small peaks in the bedrock spectra
are able to excite higher modes of vibration of the soil deposit producing a multi-peaked surface
response. This phenomenon often occurs and in some cases the peak spectral value may be
developed at a period corresponding to the predominant period of the bedrock motion. Perhaps
the answer is to use a range of possible but realistic input motions to bracket the range of input
accelerations.
The non-linear soil properties give rise to a surface motion response spectrum whose
peak spectral value and predominant period is dependent oon the magnitude of the maximum
base rock acceleration. In a truly linear system, scaling the maximum input acceleration merely
scales the peak spectral response by the same factor. The non-linearity of the soil strength
properties precludes this simple scaling as a marked change in predominant period of the surface
motion results and an analysis must be made to determine the full effect. In other words there is
no simple relationship relating maximum input acceleration to predominant period and peak
spectral velocity. Quantitatively, it is generally apparent, that deposits of deep soft soils tend to
16

produce ground motions having long period characteristics, while shallower deposits of stiff soils
tend to produce shorter period ground motions.
Clearly the level of understanding of surface layer modification of earthquake shear
waves has reached the point where engineers may now use the results of research as a useful
guide to evaluate the influence of soil condition on earthquake ground motions.
Earthquake Damage to Structures
Comparatively speaking, the absolute movement of the ground and buildings during an
earthquake is not actually all that large, even during a major earthquake. That is, they do not
usually undergo displacements that are large relative to the building's own dimensions. So, it is
not the distance that a building moves which alone causes damage.
Rather, it is because a building is suddenly forced to move very quickly that it suffers
damage during an earthquake. Think of someone pulling a rug from beneath you. If they pull it
quickly (i.e., accelerate it a great deal), then they needn't pull it very far to throw you off balance.
On the other hand, if they pull the rug slowly and only gradually increase the speed of the rug,
they can move (displace) it a great distance without that same unfortunate result.
In other words, the damage that a building suffers primarily depends not upon its
displacement, but upon acceleration. Whereas displacement is the actual distance the ground and
the building may move during an earthquake, acceleration is a measure of how quickly they
change speed as they move. During an earthquake, the speed at which both the ground and
building are moving will reach some maximum. The more quickly they reach this maximum, the
greater their acceleration.
It's worthwhile mentioning here that in order to study the earthquake responses of
buildings; many buildings in earthquake-prone regions of the world have been equipped with
strong motion accelerometers. These are special instruments which are capable of recording the
accelerations of either the ground or building, depending upon their placement.
The recording of the motion itself is known as an accelerogram. Figure below shows an
accelerogram recorded in a hospital building parking lot during the Northridge, California
earthquake of January 17, 1994.

17

In addition to providing valuable information about the characteristics of the particular


earthquake recorded or the building where the accelerogram was recorded, accelerograms
recorded in the past are also often used in the earthquake response analysis and earthquake
design of buildings yet to be constructed.
Newton's Law
Acceleration has this important influence on damage, because, as an object in movement,
the building obeys Newton' famous Second Law of Dynamics. The simplest form of the equation
which expresses the Second Law of Motion is F = MA.
This states the Force acting on the building is equal to the Mass of the building times the
Acceleration. So, as the acceleration of the ground, and in turn, of the building, increase, so does
the force which affects the building, since the mass of the building doesn't change.
Of course, the greater the force affecting a building, the more damage it will suffer;
decreasing F is an important goal of earthquake resistant design. When designing a new building,
for example, it is desirable to make it as light as possible, which means, of course, that M, and in
turn, F will be lessened. As we've seen in the discussion of Advanced Earthquake Resistant
Techniques, various techniques are now also available for reducing A.
Inertial Forces

Acceleration, Inertial Forces


It is important to note that F is actually what's known as an inertial force, that is, the force
is created by the building's tendency to remain at rest, and in its original position, even though
the ground beneath it is moving. This is in accordance with another important physical law
known as D'Alembert's Principle, which states that a mass acted upon by acceleration tends to
oppose that acceleration in an opposite direction and proportionally to the magnitude of the
acceleration.
This inertial force F imposes strains upon the building's structural elements. These
structural elements primarily include the building's beams, columns, load-bearing walls, floors,
as well as the connecting elements that tie these various structural elements together. If these
strains are large enough, the building's structural elements suffer damage of various kinds.
18

Building Frequency and Period


To begin with, as we discussed in the How Earthquakes Affect Buildings, the magnitude
of the building response that is, the accelerations which it undergoes depends primarily upon
the frequencies of the input ground motion and the building's natural frequency. When these are
near or equal to one another, the building's response reaches a peak level.
In some circumstances, this dynamic amplification effect can increase the building
acceleration to a value two times or more that of the ground acceleration at the base of the
building. Generally, buildings with higher natural frequencies, and a short natural period, tend to
suffer higher accelerations but smaller displacement. In the case of buildings with lower natural
frequencies, and a long natural period, this is reversed as the buildings will experience lower
accelerations but larger displacements.
Building Stiffness
The taller a building, the longer its natural period tends to be. But the height of a building
is also related to another important structural characteristic: the building flexibility. Taller
buildings tend to be more flexible than short buildings. (Only consider a thin metal rod. If it is
very short, it is difficulty to bend it in your hand.
If the rod is somewhat longer, and of the same diameter, it will become much easier to
bend. Buildings behave similarly.) We say that a short building is stiff, while a taller building is
flexible. (Obviously, flexibility and stiffness are really just the two sides of the same coin. If
something is stiff, it isn't flexible and vice-versa.)
Stiffness greatly affects the building's uptake of earthquake generated force. Reconsider
our first example above, of the rigid stone block deeply founded in the soil. The rigid block of
stone is very stiff; as a result it responds in a simple, dramatic manner. Real buildings, of course,
are more inherently flexible, being composed of many different parts.
Furthermore, not only is the block stiff, it is brittle; and because of this, it cracks during
the earthquake. This leads us to the next important structural characteristic affecting a building's
earthquake response and performance and ductility.
Ductility
Ductility is the ability to undergo distortion or deformation bending, for example
without resulting in complete breakage or failure. To take once again the example of the rigid
block in Figure 3, the block is an example of a structure with extremely low ductility. To see
how ductility can improve a building's performance during an earthquake, consider figure below.

19

Metal Rod Ductility


For the block, we have substituted a combination of a metal rod and a weight. In response
to the ground motion, the rod bends but does not break. (Of course, metals in general are more
ductile than materials such as stone, brick and concrete.) Obviously, it is far more desirable for a
building to sustain a limited amount of deformation than for it to suffer a complete breakage
failure.
The ductility of a structure is in fact one of the most important factors affecting its
earthquake performance. One of the primary tasks of an engineer designing a building to be
earthquake resistant is to ensure that the building will possess enough ductility to withstand the
size and types of earthquakes it is likely to experience during its lifetime.
Damping
The last of the important structural characteristics, or parameters, which we'll discuss
here is damping. As we noted earlier, ground and building motion during an earthquake has a
complex, vibratory nature. Rather than undergoing a single "yank" in one direction, the building
actually moves back and forth in many different horizontal directions
All vibrating objects, including buildings, tend to eventually stop vibrating as time goes
on. More precisely, the amplitude of vibration decays with time. Without damping, a vibrating
object would never stop vibrating, once it had been set in motion. Obviously, different objects
possess differing degrees of damping. A bean bag, for example, has high damping; a trampoline
has low damping.
In a building undergoing an earthquake, damping the decay of the amplitude of a
building's vibrations is due to internal friction and the absorption of energy by the building's
structural and nonstructural elements. All buildings possess some intrinsic damping.
The more damping a building possesses, the sooner it will stop vibrating--which of
course is highly desirable from the standpoint of earthquake performance. Today, some of the
more advanced techniques of earthquake resistant design and construction employ added
damping devices like shock absorbers to increase artificially the intrinsic damping of a building
and so improve its earthquake performance.
20

References
http://en.wikipedia.org/wiki/Earthquake
http://eqseis.geosc.psu.edu/~cammon/HTML/Classes/IntroQuakes/Notes/earthquake_size.html
http://en.wikipedia.org/wiki/Earthquake_location
http://www.isr.umd.edu/~austin/aladdin.d/matrix-eq-spectra.html
http://geology.utah.gov/utahgeo/hazards/eqfault/eqfault.htm
http://mceer.buffalo.edu/infoservice/reference_services/buildingRespondEQ.asp

21

You might also like