You are on page 1of 25

CVE 311 – Earthquake Engineering

3. Earthquake Engineering, Part 2


3.1 Duration

12 hours.

3.2 Objectives

Know the measurements of earthquake


Learn the dynamics of vibrations: attenuation
Learn about earthquake’s time history

3.3 Pre-Assessment
Before starting the proper learning of the topics, kindly answer the following. This is
not a test but is a way for us to see what you already know or do not know about the
topics.

1 2 3 4
I have no I have a I have some I know so
Topics
idea very little idea much
idea
Measurement’s of
Earthquakes
A. Intensity Scale
B. Seismographs and
Seismograms
C. Magnitude Scale
D. Intensity-Magniture
Relationships
Dynamics of
Vibrations: Attenuation
A. Earthquake
Occurrence and
Return Period
B. Ground-Motion
Models (Attenuation
Relationships)
C. Features of Strong-
Motion Data for
Attenuation
Relationship
D. Attenuation
Relationship for
Europe
E. Attenuation
Relationship for Japan
F. Attenuation
Relationship for North
America
G. Worldwide

Page 59

PDF processed with CutePDF evaluation edition www.CutePDF.com


CVE 311 – Earthquake Engineering
1 2 3 4
I have no I have a I have some I know so
Topics
idea very little idea much
idea
Attenuation
Relationship
Time History

3.4 Measurements of Earthquakes


As is well known to those who have experienced them, not all earthquakes have the
same intensity. Some are barely felt, some are felt strongly but cause only moderate
damage, and yet some others are so strong that are capable of producing
widespread and catastrophic damage.

From the engineering point of view, it is thus important to have a scale with which
one can measure or quantify the intensity of earthquakes. This chapter will describe
the different scales that throughout the years have been devised to measure the size
of earthquakes and are still of relevance today. It will also describe the instruments
that are employed nowadays to record the ground motions generated by
earthquakes and collect the information that is needed to determine the earthquake
size and the location where earthquakes originate. It will present, also, some
techniques to determine this location.

3.4.1 Intensity Scale


Intensity scales are among the first measurement systems devised to characterize
the strength of earthquakes. These scales are based on a qualitative description of
the damage caused by an earthquake to the natural and built environment at a
particular location and the associated human reaction.

The use of an intensity scale to measure the strength of an earthquake dates back to
1564 with the introduction of the Gastaldi scale.

Of a more recent vintage are the intensity scales developed in the 1880s by M. S. de
Rossi in Italy and Francois Forel in Switzerland for European conditions, and a
refined version of these scales devised by the Italian seismologist Giuseppe Mercalli
in 1902.

In recent times, the most widely used intensity scale in North America and other
parts of the world is a modified version of the Mercalli scale introduced by Harry O.
Wood and Frank Newman in 1931 for U.S. conditions. This scale, known as the
Modified Mercalli Intensity scale or MMI scale, is also based on an assessment of
the local destructiveness induced by an earthquake and the way people react to it. It
is composed of 12 grades, ranging from Grade I for an earthquake that is not felt by
the people to Grade XII for an earthquake that causes destruction.

Other modern intensity scales are the 8-grade scale of the Japanese Meteorological
Agency (JMA), developed in 1949 for Japanese conditions, and the 12-grade
Medvedev–Sponheuer–Karnik (MSK) scale, introduced in 1964 and intended for
international use.

Page 60
CVE 311 – Earthquake Engineering
3.4.1.1 Modified Mercalli Intensity (MM) Scale

I. Not felt except by a very few under especially favorable circumstances.


II. Felt only by a few persons at rest, especially on upper floors of buildings.
Delicately suspended objects may swing.
III. Felt quite noticeably indoors, especially on upper floors of buildings, but many
people do not recognize it as an earthquake. Standing motor cars may rock
slightly. Vibration like the passing of a truck. Duration estimated.
IV. During the day felt indoors by many, outdoors by few. At night some
awakened. Dishes, windows, doors disturbed; walls make cracking sound.
Sensation like heavy truck striking building. Standing motor cars rocked
noticeably.
V. Felt by nearly everyone, many awakened. Some dishes, windows, etc.,
broken; a few instances of cracked plaster; unstable objects overturned.
Disturbances of trees, poles, and other tall objects are sometimes noticed.
Pendulum clocks may stop.
VI. Felt by all, many frightened and run outdoors. Some heavy furniture moved; a
few instances of fallen plaster or damaged chimneys. Damage slight.
VII. Everybody runs outdoors. Damage negligible in buildings of good design and
construction; slight to moderate in well-built ordinary structures; considerable
in poorly built or badly designed structures; some chimneys broken. Noticed
by persons driving motor cars.
VIII. Damage slight in specially designed structures; considerable in ordinary
substantial buildings, with partial collapse; great in poorly built structures.
Panel walls are thrown out of frame structures. Fall of chimneys, factory
stacks, columns, monuments, walls. Heavy furniture overturned. Sand and
mud ejected in small amounts. Changes in well water. Persons driving motor
cars disturbed.
IX. Damage considerable in specially designed structures; well-designed frame
structures thrown out of plumb; great in substantial buildings, with partial
collapse. Buildings shifted off foundations. The ground cracked
conspicuously. Underground pipes are broken.
X. Some well-built wooden structures destroyed; most masonry and frame
structures destroyed with foundations; ground badly cracked. Rails bent.
Landslides considerable from river banks and steep slopes. Shifted sand and
mud. Water splashed over banks.

Page 61
CVE 311 – Earthquake Engineering
XI. Few, if any, (masonry) structures remain standing. Bridges destroyed. Broad
fissures in the ground. Underground pipelines completely out of service. Earth
slumps and landslips in soft ground. Rails bent greatly.
XII. Damage total. Waves are seen on the ground surface. Lines of sight and level
are distorted. Objects are thrown into the air.

3.4.1.2 Japanese Seismic Intensity (JMA) Scale


0. No sensation: registered by seismographs but no perception by the human
body

I. Slight: felt by persons at rest or persons especially sensitive to earthquakes

II. Weak: felt by most persons; the slight rattling of doors and Japanese latticed
paper sliding doors (shoji)

III. Rather strong: shaking of houses and buildings; the heavy rattling of doors
and shoji; swinging of chandeliers and other hanging objects; movement of
liquids in vessels

IV. Strong: strong shaking of houses and buildings; overturning of unstable


objects; spilling of liquids out of vessels four-fifths full

V. Very strong: cracking of plaster of walls; overturning of tombstones and stone


lanterns; damage to masonry chimneys and mud-plastered warehouses

3.4.1.3 Medvedev–Sponheuer–Karnik Intensity (MKS) Scale


I. Not noticeable. The intensity of vibration is below the limit of sensibility.
Tremor is detected and recorded by seismographs only.

II. Scarcely noticeable. Vibration is felt only by individual people at rest in


houses, especially on the upper floors of buildings.

III. Weak, partially observed only. Earthquake is felt indoors by a few people,
outdoors only in favorable circumstances. Vibration is like that due to the
passing of a light truck. Attentive observers notice a slight swinging of hanging
objects, somewhat more heavily on upper floors.

IV. Largely observed. Earthquake is felt indoors by many people, outdoors by


few. Here and there people awake, but no one is frightened. The vibration is
like that due to the passing of a heavily loaded truck. Windows, doors, and
dishes rattle. Floors and walls creak. Furniture begins to shake. Hanging
objects swing slightly. Liquids in open vessels are slightly disturbed. In
standing motor cars the shock is noticeable.

V. Awakening. Earthquake is felt indoors by all, outdoors by many. Many


sleeping people awake. A few run outdoors. Animals become uneasy.
Buildings tremble throughout. Hanging objects swing considerably. Pictures
knock against walls or swing out of place. Occasionally pendulum clocks stop.
Few unstable objects may be overturned or shifted. Open doors and windows
are thrust open and slam back again. Liquids spill in small amounts from well-
Page 62
CVE 311 – Earthquake Engineering
filled open containers. The sensation of vibration is like that due to a heavy
object falling inside the building. Slight damage of Grade1 1 in buildings of
Type 2 A is possible. Sometimes change in the flow of springs.

VI. Frightening. Felt by most indoors and outdoors. Many people in buildings are
frightened and run outdoors. A few persons lose their balance. Domestic
animals run out of their stalls. In a few instances, dishes and glassware may
break, books fall. Heavy furniture may move and small steeple bells may ring.
Damage of Grade 1 is sustained in single buildings of Type B and many of
Type A. Damage in few buildings of Type A is of Grade 2. In few cases cracks
up to widths of 1 cm possible in wet ground. In mountains, occasional
landslips; changes in the flow of springs, and level of well water are observed.

VII. Damage to buildings. Most people are frightened and run outdoors. Many find
it difficult to stand. The vibration is noticed by persons driving motor cars.
Large bells ring. In many buildings of Type C damage of Grade 1 is caused; in
many buildings of Type B damage is of Grade 2. Many buildings of Type A
suffer damage of Grade 3, few of Grade 4. In single instances landslips of
roadway on steep slopes; cracks in roads; seams of pipelines damaged;
cracks in stone walls. Waves are formed on the water, and water is made
turbid by mud stirred up. Water levels in wells change, and the flow of springs
changes. In few cases dry springs have their flow restored and existing
springs stop flowing. In isolated instances, parts of sandy or gravelly banks
slip off.

VIII. Destruction of buildings. Fright and panic; also persons driving motor cars are
disturbed. Here and their branches of trees break off. Even heavy furniture
moves and partly overturns. Hanging lamps are in part damaged. Many
buildings of Type C suffer damage of Grade 2, few of Grade 3. Many buildings
of Type B suffer damage of Grade 3 and a few of Grade 4, and many
buildings of Type A suffer damage of Grade 4 and a few of Grade 5.
Occasional breakage of pipe seams. Memorials and monuments move and
twist. Tombstones overturn. Stone walls collapse. Small landslips in hollows
and on banked roads on steep slopes; cracks in ground up to widths of
several centimeters. Water in lakes becomes turbid. Dry wells refill and
existing wells become dry. In many cases change in flow and level of water.

IX. General damage to buildings. General panic; considerable damage to


furniture. Animals run to and fro in confusion and cry. Many buildings of Type
C suffer damage of Grade 3, a few of Grade 4. Many buildings of Type B
show damage of Grade 4, a few of Grade 5. Monuments and columns fall.
Considerable damage to reservoirs; underground pipes partly broken. In
individual cases, railway lines are bent and roadways are damaged. On flat
land overflow of water, sand, and mud is often observed. Ground cracks to
widths of up to 10 cm, on slopes and river banks more than 10 cm;
furthermore a large number of slight cracks in the ground; fall of rocks, many
landslides, and earth flows; large waves on water. Dry wells renew their flow
and existing wells dry up.

X. General destruction of buildings. Many buildings of Type C suffer damage of


Grade 4, a few of Grade 5. Many buildings of Type B show damage of Grade
5; most of Type A have destruction Type 5; critical damage to dams and dikes
and severe damage to bridges. Railway lines are bent slightly. Underground
pipes are broken or bent. Road paving and asphalt show waves. In-ground,
Page 63
CVE 311 – Earthquake Engineering
cracks up to widths of several decimeters, sometimes up to 1 m. Parallel to
watercourses broad fissures occur. Loose ground slides from steep slopes.
From river banks and steep coasts, considerable landslides are possible. In
coastal areas displacement of sand and mud; change of water level in wells;
water from canals, lakes, rivers, etc., thrown on land. New lakes occur.

XI. Catastrophe. Severe damage even to well-built buildings, bridges, water


dams, and railway lines; highways become useless; underground pipes
destroyed. Ground considerably distorted by broad cracks and fissures, as
well as by movement in horizontal and vertical directions; numerous landslips
and falls of rock.

XII. Landscape changes. Practically all structures above and below ground are
greatly damaged or destroyed. The surface of the ground is radically
changed. Considerable ground cracks with extensive vertical and horizontal
movements are observed. Falls of rock and slumping of river banks over wide
areas; lakes are dammed; waterfalls appear, and rivers are deflected.

Damage to buildings is classified into the following five grades:

Grade 1: Slight damage: Fine cracks in plaster; fall of small pieces of


plaster
Grade 2: Moderate damage: Small cracks in walls; fall of fairly large pieces of
plaster; particles slip off; cracks in chimneys; parts of chimneys fall
down
Grade 3: Heavy damage: Large and deep cracks in walls; fall of chimneys
Grade 4: Destruction: Gaps in walls; parts of buildings may collapse;
separate parts of building lose their cohesion; inner walls and filled-
in walls of the frame collapse
Grade 5: Total damage: Total collapse of buildings

Buildings are classified into the following three types:


Structure A: Buildings in fieldstone, rural structures, adobe houses, clay
houses
Structure B: Ordinary brick buildings, buildings of the large block and
prefabricated type, half-timbered structures, buildings in natural
hewn stone
Structure C: Reinforced buildings, well-built wooden structures

In dealing with intensity scales, it is important to keep in mind that they do not involve
a precise scientific measurement of the severity of earthquakes and are therefore of
limited value. The problem with these scales is that they depend on subjective
factors such as:

a. previous experience of people with earthquakes,


b. local design and construction practices,
c. whether or not the earthquake occurs in an inhabited region, and
d. the population density.

For example, the description many frightened in the MMI scale will depend on the
location of the earthquake. A tremor that would alarm the residents of Cleveland,
Ohio, would most likely be ignored by people in Tokyo or Los Angeles. Likewise, the
Page 64
CVE 311 – Earthquake Engineering
collapse of buildings, a key factor in determining an intensity rating, may not only
reflect the power of an earthquake but also whether or not the collapsed buildings
were designed to resist seismic loads. Intensity scales cannot therefore by their
nature be accurate.

Despite their limitations, intensity scales may be useful to estimate the size and
location of earthquakes that occurred before the development of modern seismic
instruments. Because qualitative descriptions of the effects of earthquakes are often
available through historical records, intensity scales may be used to characterize the
rate of earthquake recurrence at the locations wherever these historical records are
available.

Intensity scales may also be useful to describe the distribution of damage in a


region, to identify areas of poor soils, and to approximately locate the earthquake
epicenter. For this reason, even in modern times, contours of equal intensities, or
isoseismal, are routinely plotted over a map of the geographical regions where
strong earthquakes occur. Examples of these plots, called isoseismal maps, are
shown below:

Page 65
CVE 311 – Earthquake Engineering

Some of the most common intensity scales are listed below:

i. Mercalli–Cancani–Seiberg (MCS): 12‐level scale used in Southern Europe.


ii. Modified Mercalli (MM): 12‐level scale proposed in 1931 by Wood and
Neumann, who adapted the MCS scale to the California data set. It is used in
North America and several other countries.
iii. Medvedev–Sponheuer–Karnik (MSK): 12‐level scale developed in Central
and Eastern Europe and used in several other countries.
iv. European Macroseismic Scale (EMS): 12‐level scale adopted since 1998 in
Europe. It is a development of the MM scale.
v. Japanese Meteorological Agency (JMA): 7‐level scale used in Japan. It has
been revised over the years and has recently been correlated to the maximum
horizontal acceleration of the ground.

3.4.2 Seismographs and Seismograms


Aware that intensity scales were based on subjective appraisals of damage
submitted by a wide variety of observers, scientists studying earthquakes at the end
of the nineteenth century realized that an understanding of the earthquake
phenomenon required accurate and consistent physical measurements. They felt,
thus, the need for advanced devices that would record and preserve the ground
motion generated by earthquakes.

Page 66
CVE 311 – Earthquake Engineering
One of the first of such devices was developed by Filippo Cecchi in Italy in 1875, it
was designed to start a clock and a recording device at the first sign of shaking. It
then recorded the relative motion between a pendulum and the shaking ground as a
function of time. The oldest known record produced by this instrument is dated
February 23, 1887.

Afterward, John Milne, an English geologist working in Japan in the late 1800s,
developed with the help of James Ewing, a professor of mechanical engineering and
physics, and Thomas Gray, a professor of telegraphic engineering, the first
instrument to record the motion of the ground in all of its three directions: up and
down, back and forth, and side to side. This was accomplished with three
independent pendulums, each with an attached stylus that inscribed the motion of
the pendulum on a roll of smoked paper. Also, the instrument was implemented with
a clock mechanism that enabled each device to operate for 24 h at a time and stamp
the paper with the precise time of the first wave arrival. This instrument became
known as the Milne seismograph and was for many years the standard equipment
for seismologists around the world. With it, a new era in the study of earthquakes
began.

The seismograph that Milne, Ewing, and Gray developed at the end of the
nineteenth century fell short of meeting the demands of a science that had more
questions than answers. One of its shortcomings was that it responded only to
motions with periods within a narrow range. Another was the tendency of the
seismograph’s pendulum to keep swinging indefinitely once in motion. Without a way
to control this free-vibration motion, the seismograph was unable to record
accurately the motion generated by late-arrival waves. A substantial improvement
came in 1898, when Emil Wiechert in Germany introduced a viscous damping
mechanism that restrained the seismograph.

The next advance in the development of the modern seismograph was made in 1914
by Boris Galitzin, a Russian seismologist. Galitzin introduced a design that did away
with the need for a mechanical linkage between pendulum and recorder. He
mounted a wire coil on the seismograph’s pendulum and suspended the coil and the
pendulum between the poles of a magnet fixed to the ground. This way, the motion
of the pendulum generated an electric current that was proportional to the
pendulum’s velocity, which Galitzin used to rotate a galvanometer coil. Light
reflected from a mirror on the galvanometer coil was then recorded on photographic
paper. This was a development that dominated the seismograph design throughout
the twentieth century and had profound implications for seismology. With it,
designers were able to use smaller pendulum masses and eliminate the unwanted
friction between a stylus and a recording medium. Equally important, it meant the
ability to transmit the recordings from remote seismographs to a central location,
eliminating thus the need for traveling periodically from site to site to monitor the
instruments.pendulum and greatly improved its accuracy.

3.4.2.1 COMPONENTS AND DESIGN FEATURES


Shown below is the schematic diagram of a seismograph; (a) horizontal (b) vertical.
They are ordinarily composed of a pendulum, a damping element, a stylus attached
to the pendulum mass, and a recording drum.

Page 67
CVE 311 – Earthquake Engineering

To be able to detect distant earthquakes, seismographs are also implemented with a


mechanical, optical, or electromagnetic system that magnifies the movement of the
pendulum up to a hundred-, thousand-, or million-fold, respectively. The damping
element is normally an air, oil, or electromagnetic damper and is introduced into the
system to damp out the motion of the pendulum as soon as the ground motion stops.
Dampers with a damping ratio of the order of 80% of critical are used for this
purpose. The recording drum holds wrapped around a waxed or smoked paper on
which the stylus inscribes the movement of the pendulum. In some designs, this
motion is recorded on photographic paper or magnetic tape. The recording drum
continuously rotates at a fixed rate and moves from left to right on its shaft in a
period of 24 h, producing thus a continuous record of the pendulum motion.

However, when the paper is removed from the drum and is laid flat, the record
appears as several parallel lines, as it can be seen in the typical record shown
below.

Another design feature of seismographs is the time marks they introduce on the
record at regular intervals. These time marks are produced by deflecting the trace for
a second or two, usually at the end of every minute (see above figure). By
convention, seismographs are marked in terms of Universal Time Coordinated (UTC)
[also called Greenwich Mean Time (GMT)], not local time. Seismographs operate 24
h a day, 365 days a year. As earthquake-generated ground motions are neither
purely vertical nor purely horizontal, in practice a vertical seismograph and two
horizontal seismographs are used to record three components of ground motion
along three orthogonal directions. This way the motion of the ground at any instant
may be completely characterized by the vector sum of the motion recorded along
each of these three directions. The below figure shows the interior of a modern
electromagnetic seismograph.

Page 68
CVE 311 – Earthquake Engineering

To gain, then, an insight as to how a seismograph measures the motion of the


ground, consider the following simplified formulation. Assume first that the
earthquake ground motion is predominantly sinusoidal with amplitude u0 and
dominant period Tg, and that the displacement, velocity, and acceleration of the
ground may be expressed approximately as:

2 2 2
sin ; sin ; sin

where:

2 2
;

Consider then that in terms of such assumptions the relative motion between the
pendulum and the ground may be expressed by the differential equation of motion

2 2
sin

where m denotes the mass of the pendulum, c is the damping coefficient of the
seismograph’s damper, k the corresponding spring constant, and u the displacement
of the pendulum relative to the ground surface. Consider, also, that the solution of
the above equation is of the form:
2
sin !

"#1 ! & #2' ! &


% %

where Tn and ξ, respectively, denote the seismograph’s natural period and damping
ratio, and θ is a phase angle. Hence, the displacement amplitude recorded by a
seismograph divided over that of the ground may be expressed as:

Page 69
CVE 311 – Earthquake Engineering
1

"#1 ! & #2' ! &


% %

3.4.2.2 SEISMOGRAMS
The records obtained from a seismograph are called seismograms. A seismogram is
thus a record of the variation with time of the displacement of the ground, magnified
by the magnification factor of the seismograph, at the location where the
seismograph is installed. A typical seismogram is shown below. The numbers in the
middle of the record indicate the hours referred to the GMT. The small deflections at
regular intervals along the trace are time marks at 1 min intervals. There are 60 such
marks in each line, so each line represents the motion recorded during 1 h.

It should be noted that the traces in a seismogram are never without some little
ripples. These little ripples show up in a seismogram because seismographs are so
sensitive that they can detect the ever-present background noise of the earth. They
are called microseisms and arise from local disturbances such as traffic on the
streets, the effect of winds on trees, breaking of the surf on the beach, and other
natural and human-made disturbances.

3.4.3 Magnitude Scale

3.4.3.1 RICHTER OR LOCAL MAGNITUDE


Besides providing information for the location of earthquakes, seismograms also
provide the information that is needed to estimate the size or strength of an
earthquake in terms of what is called earthquake magnitude. This instrumentally
quantified measure of earthquake strength is widely used nowadays by
seismologists, engineers, and even the general public. Although in some cases it
fails to give an accurate representation of the true strength of an earthquake, it is still
routinely used to characterize the intensity of earthquakes and remains a key
parameter in earthquake hazard analysis.

The concept of earthquake magnitude was introduced by Charles Richter in 1935 to


overcome the limitations of the intensity scales, the only method used back then to
describe and compare earthquakes. Following a fundamental idea first used by K.
Page 70
CVE 311 – Earthquake Engineering
Wadati in Japan, Richter based his magnitude scale on a measurement of the wave
motion recorded by a seismograph. He borrowed the term magnitude from
astronomy as the relative brightness of stars (stellar magnitude) is referred to as
magnitude. However, the analogy stops there because in astronomy a smaller
magnitude means an increased brightness.

Richter defined his scale in terms of the peak amplitude of the trace recorded by the
then-standard Wood–Anderson seismograph, which, as observed earlier, has a
magnification factor of 2800, a natural period of 0.8 s, and a damping ratio of 80%.
However, because such an amplitude can vary significantly from earthquake to
earthquake, he used the logarithm of it, as opposed to the amplitude itself, to
compress the range of the scale. Similarly, as the amplitude of seismic waves
decreases with distance from the earthquake epicenter, he set the measurement of
this amplitude at a standard distance of 100 km. Furthermore, he described such a
peak trace amplitude about the peak trace amplitude that would be generated by a
zero magnitude earthquake; that is, a barely perceptible earthquake. For this
purpose, he defined a zero-magnitude earthquake as that which theoretically would
produce a seismogram with a peak trace of 1 µm (10−6 m) at a distance of 100 km.
As introduced by Richter, earthquake magnitude is thus defined as the logarithm to
base ten of the peak wave amplitude measured in micrometers recorded by a
Wood–Anderson seismograph at a distance of 100 km from the earthquake
epicenter. That is,
,
( log
,-

where:

A is the peak amplitude in micrometers in a seismogram (magnified ground


displacement) recorded at 100 km from the earthquake epicenter

A0 the peak amplitude of a zero-magnitude earthquake, defined as 1 µm at


a distance of 100 km.

Example 1 - Calculation of earthquake magnitude

The seismogram from an earthquake recorded by a seismograph located


exactly at 100 km from the earthquake epicenter exhibits a peak amplitude of 1
mm. Determine the magnitude of the earthquake.

Solution: A = 1 mm (1000 µm / 1mm) = 1000 µm A0 = 1 µm


, 1000
( log log 3
,- 1

Example 2- Ground displacement induced by the magnitude 9.0 earthquake

Determine the peak ground displacement induced by an earthquake with a


magnitude of 9.0 at a distance of 100 km from the earthquake epicenter.

Solution: M=9 A0 = 1 µm
,
9 log
1

Page 71
CVE 311 – Earthquake Engineering
log , 9

, 100 μm 1000m

Hence, as the magnification factor of a Wood–Anderson seismograph is 2800, an


earthquake of magnitude 9.0 induces a ground displacement of 1000m/2800 = 0.38
m at a distance of 100 km from its epicenter.

As no seismograph is likely to be located exactly at 100 km from an earthquake’s


epicenter, an extrapolation or correction is needed to be able to determine
earthquake magnitudes from seismograms obtained at epicentral distances other
than 100 km. A common procedure is that of constructing a curve that defines the
variation of ground motion amplitude with distance for the zero-magnitude
earthquake. This curve, known as the A0 curve, is determined empirically using data
from a large number of earthquakes and is unique for each seismological station.
They reflect the geological and geophysical conditions that surround each station.
Using this empirical curve and adopting the assumption that the ratio of the peak
ground motion amplitudes at two given distances is the same independently of
earthquakes magnitude, it is then possible to compute the magnitude of an
earthquake using the original definition and data from any seismogram. To do so,
one simply considers that A represents the peak amplitude in the selected
seismogram and A0 the zero-magnitude amplitude that corresponds to the epicentral
distance to the station from which the seismogram is obtained. Richter, for example,
developed the following empirical equation to define the variation with distance of the
amplitude A0 for earthquakes in Southern California:

log , 5.12 2.56 log Δ

In this equation, A0 is in micrometers, ∆ epicentral distance in kilometers, and 10 < ∆


< 600 km.

Problem 3 - Earthquake magnitude in terms of A0 empirical equation

Determine the magnitude of an earthquake recorded in Southern California at a


seismological station located 147 km from the earthquake epicenter. The
recorded peak amplitude was 66 mm.

Solution: ∆ = 147 km A = 66 mm = 66,000 µm

log , 5.12 2.56 log Δ 5.12 2.56 log 147

, 109.: ; .9< =>? :@A


0.37 μm

Hence, the magnitude of the earthquake is equal to


, 66000
( log log 5.25
,- 0.37

3.4.3.2 SURFACE- AND BODY-WAVE MAGNITUDES


Because Richter’s original scale is limited to local earthquakes at epicentral
distances of no more than 600 km and recorded in only one kind of instrument, the
desire for the global characterization of earthquake size made it necessary for the
introduction of a new definition of earthquake magnitude. This new definition was the
Page 72
CVE 311 – Earthquake Engineering
surface-wave magnitude introduced by B. Gutenberg and Richter himself in 1936.
The definition of this magnitude is the same as that originally introduced by Richter,

wave with a period of ∼20 s. Beyond ∼600 km, the seismograms of shallow
except that the measured amplitude corresponds, by convention, to that of a surface

with a period of ∼20 s. In general, therefore, surface-wave magnitude is determined


earthquakes recorded by long-period seismographs are dominated by surface waves

using seismograms from long-period seismographs. In general, too, the record from
the vertical component of motion is used in this definition. The symbol used to
identify surface-wave magnitude is Ms. Using the format of Richter’s equations,
surface-wave magnitude is at times alternatively defined by:

(B log 1.66 log Δ 2

where is the peak ground displacement measured in micrometers and Δ epicentral


distance measured in degrees (360° corresponding to the circumference of the
earth).

Problem 4 - Calculation of surface-wave magnitude

The peak amplitude of the Rayleigh wave (reduced to ground displacement) in


a seismogram from a seismographic station located at a distance of 28° from
an earthquake’s epicenter is 4.3 µm. Determine using this information the
earthquake’s surface-wave magnitude.

Solution: = 4.3 µm Δ = 28°

(B log 1.66 log Δ 2 log 4.3 1.66 log 28 2 5.04

3.4.3.3 Body Wave Magnitude


Body wave magnitude (mb): measures the amplitude of P‐waves with a period of
about 1.0 second, which is less than 10 km wavelengths. This scale is suitable for
deep earthquakes that have few surface waves. Moreover, mb can measure distant
events, for example, epicentral distances not less than 600 km. Furthermore, P‐
waves are not affected by the depth of the energy source. Magnitude mb is related to
the amplitude A and period T of P‐waves as follows:
,
D log ! E log Δ

in which σ(∆) is a function of the epicenter distance ∆ (in degrees). For example, if ∆
= 45° then σ = 6.80; other values can be found in the literature (e.g. Udias, 1999).

3.4.3.4 Moment Magnitude


Moment magnitude (MW ): accounts for the mechanism of shear that takes place at
earthquake sources. It is not related to any wavelength. As a result, Mw can be used
to measure the whole spectrum of ground motions. Moment magnitude is defined as
a function of the seismic moment M0. This measures the extent of deformation at the
earthquake source and can be evaluated as follows:

( F , Δu

in which G is the shear modulus of the material surrounding the fault, A is the fault
rupture area and ∆u is the average slip between opposite sides of the fault. The

Page 73
CVE 311 – Earthquake Engineering
modulus G can be assumed to be 32 000 MPa in the crust and 75 000 MPa in the
mantle. Mw is thus given by:

(H 0.67 log ( 10.70

where M0 is expressed in ergons.

3.4.4 Intensity–Magnitude Relationships


Intensity–magnitude relationships are essential for the use of historical earthquakes
for which no instrumental records exist. Several simple methods to convert intensity
into magnitude have been proposed (e.g. Lee et al., 2003); most of which exhibit
large scatter because of the inevitable bias present in the definition of intensity
(Ambraseys and Melville, 1982). Gutenberg and Richter (1956) proposed a linear
relationship between local magnitude ML and epicentral intensity I0 for southern
California, given by:

(I 0.67J 1.0

in which the intensity I0 is expressed in the MM scale. The above equation shows, for
example, that the epicentral intensity I0 of VI corresponds to ML = 5.02 indicating that
the earthquake is likely to cause significant damage.

Street and Turcotte (1977) related mb magnitude to the intensity I0 (in the MM scale)
as follows:

D 0.49J 1.66

which is useful in converting earthquake data in the central and eastern USA. The
above equation relates to an intensity of VI in the MM scale to a magnitude mb of
4.60, which contradicts the observation that ML should be systematically lower than
mb for short‐period waves.

Intensity–magnitude relationships were proposed by Ambraseys (1985, 1989) for


European regions as follows:

(K 1.10 0.62JL 1.30 ∗ 10;N OL 1.62 log OL

which is applicable for northwest Europe, and

(K 0.90 0.58JL 1.10 ∗ 10;N OL 2.11 log OL

for the Alpine zone, where Ii is the MM intensity of the ith isoseismal and ri is the
radius of the equivalent area enclosed by the ith isoseismal, in kilometers.

Local geological conditions and focal depths can significantly affect the intensity of
earthquake ground motion. Semi‐empirical formulations accounting for focal depths
are available (e.g. Kanai, 1983). Sponheuer (1960) proposed to calculate M from the
epicentral intensity I0 as follows:

(K 0.66J 1.7 log ℎ 1.4

where the focal depth h is in kilometers and the intensity I0 is in the MM scale.

Page 74
CVE 311 – Earthquake Engineering
Attenuation relationships (relationships between a ground‐shaking parameter,
magnitude, distance, and soil condition) for different ground‐motion parameters can
be derived from intensity and magnitude; they may account for distance, travel path,
and site effects.

3.5 Dynamics of Vibrations: Attenuation


Earthquake response of structures and their foundations is an outcome of the
complex interaction between the random input ground motion and the continuously
changing dynamic characteristics of the system subjected to the ground motion.
Therefore, to arrive at a reliable assessment of assets, a complete understanding of
both input motion and structural system, and their interaction, is required.

3.5.1 Earthquake Occurrence and Return Period


It is of importance to estimate the frequency of occurrence of earthquakes that are
likely to occur in an area that may influence the construction site during the lifetime
of the intended facility. Account should be taken of the uncertainty in the demand
imposed by the earthquake, as well as the uncertainty in the capacity of the
constructed facility. Current seismic design approaches deal with uncertainties
associated with structural demand and capacity by utilizing probabilistic analysis.

Earthquakes are usually modeled in probabilistic seismic hazard assessment as a


Poisson process. The Poisson model is a continuous-time, integer‐value counting
process with stationary independent increments.

Earthquake ground‐motion representations for seismic structural assessment.

3.5.2 Ground‐Motion Models (Attenuation Relationships)


The ‘attenuation’ of earthquake ground motions is an important consideration in
estimating ground‐motion parameters for assessment and design purposes. Ground‐
motion models (or attenuation relationships) are analytical expressions describing
ground‐motion variation with magnitude, source distance, and site condition, which

Page 75
CVE 311 – Earthquake Engineering
account for the mechanisms of energy loss of seismic waves during their travel
through a path. Attenuation relationships permit the estimation of both the ground
motion at a site from a specified event and the uncertainty associated with the
prediction. This estimation is a key step in probabilistic and deterministic seismic
hazard analysis (Cornell, 1968). Several ground‐motion models have been
developed by various researchers. Relationships based on peak ground‐motion
parameters (PGA, peak ground velocity, PGV and peak ground displacement, PGD)
and spectral acceleration, velocity, and displacement parameters (Sa, Sv, and Sd),
are generally employed in structural earthquake engineering.

Empirical approaches generally match the data to a functional form derived from the
theory; in turn, theoretical approaches often use empirical data to determine the
values of parameters. The functional form for ground‐motion attenuation
relationships is as follows:

log Q log R: logST: ( U logST V U logSTN (, V U logST@ XL U log Y

where Y is the ground motion parameter to be computed, for example, PGA, PGV,
PGD, Sa, Sv or Sd, and b1 is a scaling factor. The second‐to‐fourth terms on the right‐
hand side are functions fi of the magnitude M, source‐to‐site distance R, and
possible source, site, and/or geological and geotechnical structure effects Ei.
Uncertainty and errors are represented by the parameter ε.

Attenuation of peak ground horizontal acceleration: effect of magnitude (a) and focal
depth (b).

Closed‐form relationships between PGA and relevant intensity scales have been
established in Japan and the USA. These are given by Kanai (1983) as follows:

ZF, 0.25 ∗ 10 .9[\]^

ZF, 0.91 ∗ 10 .N:[]]

Page 76
CVE 311 – Earthquake Engineering
in which IJMA and IMM are the values of intensity in the Japanese Meteorological
Agency (JMA) and Modified Mercalli (MM) scales, respectively. In the above
equations, the values of PGA are expressed in cm/s2.

Similarly, Trifunac and Brady (1975) suggested the following relationships for
horizontal peak ground acceleration (HPGA) and velocity:

ZF, 1.02 ∗ 10 .N[]]

ZF_ 0.23 ∗ 10 . 9[]]

where the values of PGA and PGV are in cm/s2 and cm/s, respectively.

Problem 1

Modified Mercalli intensity IMM of IX was assigned to an area of about 80 km


long and 30 km wide during an earthquake that occurred in the western United
States. Compute the peak ground acceleration (PGA) from this earthquake.
Compute the value of PGA by using both equations. Estimate the intensity IJMA
of the earthquake in the Japanese Meteorological Agency (JMA) scale.

Solution:

ZF, 0.91 ∗ 10 .N:[]]


0.91 ∗ 10 .N: 0
561.1 cm/s

ZF, 1.02 ∗ 10 .N[]]


1.02 ∗ 10 .N 0
511.2 cm/s

Estimation for IJMA: using PGA of 511.2 cm/s2

ZF, 0.25 ∗ 10 .9[\]^

511.2 0.25 ∗ 10 .9[\]^

511.2
10 .9[\]^
0.25
511.2
ln ln 10 .9[\]^
0.25
ln 2044.8 0.5Jbcd ln 10

ln 2044.8
Jbcd 6.62 ≈ 7
0.5 ln 10
3.5.3 Features of Strong‐Motion Data for Attenuation Relationships
The strong‐motion data set (or catalog) used for attenuation relationship derivation
has to fulfill several requirements. First, all magnitudes should be uniformly
recalculated using consistent approaches. Second, all distances have to be defined
uniformly. It is necessary to use the distance from the closest point on the causative
fault to the measuring site, not the epicentral distance. This is particularly important
when considering large magnitude earthquakes at short‐to‐medium distances.

Calculation of the above‐mentioned distance is an involved task that requires deep


knowledge of the local tectonic setting especially when there is no surface
Page 77
CVE 311 – Earthquake Engineering
manifestation of the fault. Moreover, the data set should be well populated and
reasonably represent distributions in magnitude, distance, and soil condition;
otherwise the ensuing attenuation relationship will exhibit statistical bias. Strong‐
motion records in databanks may have errors due to instruments and/or digitization.
Since the short-and long‐period errors present in each record are unique for each
type of instrument and digitization procedure, and because of the random nature of
the errors, each accelerogram should ideally be corrected individually. Records from
analog instruments are particularly affected by long‐period (or low‐frequency, e.g.
less than 0.5 Hz) errors because of the digitization stage which is not required for
records from digital instruments. Low‐frequency errors affect the PGV as well as the
corresponding spectral values.

3.5.4 Attenuation Relationship for Europe


Comprehensive and systematic seismological studies in Europe aimed at defining
ground-motion models for seismic hazard assessment and structural engineering
applications were conducted by Ambraseys (1975). A great deal of research has
been conducted since then at Imperial College, London, and attenuation
relationships have been formulated for Europe and the Middle East.

Revised attenuation relationships for European countries and some regions in the
Middle East have been formulated for both HPGA by Ambraseys et al. (2005a) and
for vertical peak ground acceleration (VPGA) by Ambraseys et al. (2005b). The
ground‐motion model for the HPGA is given by:

log ZF, f 2.522 0.142(H 0.314(H 3.184 log g57.76 h 0.137iK


0.05id 0.084jk 0.062jl 0.044jm

with PGA expressed in m/s2 and d is the distance (in kilometers) to the projection of
the fault plane on the surface. The latter does not require a depth estimate, generally
associated with large errors.

The coefficients SA and SS are obtained from the below table:

The coefficients FN, FO, and FT are obtained from the below table:

Page 78
CVE 311 – Earthquake Engineering
The standard deviations σ for the equation above depend on the earthquake
magnitude Mw:

(intra-plate) σ1 = 0.665 – 0.065 Mw

(inter-plate) σ2 = 0.222 – 0.022 Mw

To obtain a viable distribution of records at all magnitudes, records from earthquakes


with Mw < 5 were not considered. This also excludes records from small earthquakes
that are unlikely to be of engineering significance. The data set includes records with
magnitude Mw ranging between 5.0 and 7.6, with distances d < 100 km. Therefore,
the possible bias due to non‐triggering instruments and the effects of anelastic decay
in different regions were reduced. Moreover, most ground motions were obtained
from free‐field stations although some were recorded from either basements or
ground floors of relatively light structures that are unlikely to modify the motion from
that of the free field.

3.5.5 Attenuation Relationship for Japan


Several studies have attempted to define analytical models for the ground‐motion
parameters in Japan (e.g. Iwasaki et al., 1980; Kawashima et al., 1986; Fukushima
et al., 1995; Kamiyama, 1995). Some have also concentrated on specific areas of
the country, such as the Kanto region (e.g. Tong and Katayama, 1988). Takahashi et
al. (2000) proposed the following attenuation relationship for Japan:

log ZF, 0.446(H 0.00350h log h 0.012 ∗ 10 .@@<cn


0.00665 ℎ 20 i

where PGA is given in cm/s2. The terms d and h are focal distance and depth (in
kilometers), respectively. S is a coefficient depending on the soil type (rock, hard,
medium, and soft medium were considered in the regression analyses, as given in
the below table. However, in many circumstances, the site conditions of records
used were either unknown or uncertain. In such cases, the mean site term S can be
assumed equal to 0.941. MW is the moment magnitude.

Model errors σ were computed as follows:

E oE: E:

where σ1 and σ2 are residuals for inter-and intra‐plate earthquakes, respectively;


values decrease with increasing magnitude. It may be assumed that the total scatter
σ is equal to 0.24.

Page 79
CVE 311 – Earthquake Engineering
3.5.6 Attenuation Relationships for North America

3.5.6.1 Central and Eastern United States


The Mid‐America Earthquake Center developed a ground‐motion model to predict
HPGA in the central and eastern United States (CEUS) region (Fernandez and Rix,
2006). The attenuation relationship is based on a stochastic method and employs
three source models, that is Atkinson and Boore (1995), Frankel et al. (1996), and
Silva et al. (2003). It was developed for soil sites in the Upper Mississippi
Embayment in the New Madrid seismic zone, which is a low probability–high impact
source of earthquakes. The region has been hit by three great earthquakes in 1811
and 1812 (e.g. Reiter, 1990).

The ground‐motion model is defined by the following equations:


V
ln ZF, (H (H 6 ln Vc p qln ! , 0r < Vc
: N @ 9
70
where the equivalent distance term RM is given by:

Vc V A exp v (H

and the logarithmic standard deviation of PGA, termed σln(PGA), is considered to be


magnitude dependent. It is obtained from the following equation:

E=w xyd 0 (H :

In the above equations, R is the epicentral distance (in kilometers), Mw is the


moment magnitude, and c1 through c10 are the regression coefficients. The value of
the PGA is expressed in g. The epicentral distance R is the distance from the
observation point to the surface projection of the hypocentre. The ci coefficients,
which depend on the source model, the stress drop, the soil profile, dynamic soil
properties, and depth, can be found in Fernandez (2007). These coefficients are
computed for epicentral distances uniformly distributed between 1 and 750 km for
eight values of magnitudes Mw varying between 4 and 7.5. The equivalent distance
term RM accounts for the increase in traveling distance by the seismic waves due to
the increase in fault rupture size. The exponential term accounts for the magnitude‐
dependence of the energy release. The effects of inelastic soil behavior are
incorporated in the above attenuation relationship.

3.5.6.2 Western North America


Boore et al. (1997) and Boore (2005) formulated the following equation to predict
PGAs in western North America:

log ZF, R: R (H 6 RN (H 6 R@ V R9 log V Rz log K log d

in which R is the focal distance given by:

V gh ℎ

where d and h are the epicentral distance and the focal depth, respectively; they are
both expressed in kilometers. The value of the PGA is expressed in g.

Shallow earthquakes are those for which the fault rupture has a depth of 20 km or
less. Mw is the moment magnitude. Coefficients b1 through b5 depend on the

Page 80
CVE 311 – Earthquake Engineering
component of ground motion used. For randomly oriented horizontal components,
the PGA is given by:

log ZF, 0.105 0.229 (H 6 0.778 log V 0.371 log K log d

and the focal depth h should be assumed equal to 5.57 km; the value of vA is 1400
m/s. Thus, the resulting scatter σ is 0.160. On the other hand, for larger horizontal
components, then the equation should be modified as follows:

log ZF, 0.038 0.216 (H 6 0.777 log V 0.364 log K log d

in which the focal depth h should be 5.48 km, the value of vA is 1390 m/s and the
resulting scatter is 0.144. Site conditions are accounted for by the average shear
wave velocity to a depth of 30 m (vS, in m/s). Three soil types were considered in the
study; values for vS,30 are summarised below.

It is worth mentioning that in the derivation of the above attenuation relationships,


most of the earthquake ground motions were recorded at epicentral distances less
than 80 km, thus extrapolations should be assessed carefully based on engineering
judgment.

3.5.7 Worldwide Attenuation Relationships


Attempts to provide ground‐motion models applicable worldwide were initiated in the
1980s (Aptikaev and Kopnichev, 1980; Campbell, 1985) and continued during the
1990s (Campbell, 1993, 1997; Sarma and Srbulov, 1996, 1998). In some cases, the
attenuation relationships were derived for specific fault rupture mechanisms, such as
subduction zones (Youngs et al., 1988; Crouse, 1991; Youngs et al., 1997) or
extensional regimes (Spudich et al., 1997, 1999). Formulae for intra‐plate regions
have also been proposed by Dahle et al. (1990). Comprehensive analytical studies
based on large data sets of records for both horizontal and vertical components have
been carried out by Bozorgnia et al. (2000), Campbell and Bozorgnia (2003), and
Ambraseys and Douglas (2003); the latter is presented herein.

The form of the equation to predict HPGAs PGAh is as follows:

log ZF, f 0.659 0.202(K 0.0238h 0.020id 0.029iK

in which the epicentral distance d is in kilometers; the scatter σ is 0.214. Coefficients


SA and SS account for the effects of soil condition. Four soil categories were
considered (rock, stiff soil, soft and very soft soil); they are classified based on the
average shear wave velocity to 30 m depth vS,30. Values can be obtained from the
below table. Focal depths h are not greater than 20 km (1 ≤ h ≤ 19 km). The value of
PGA is expressed in m/s2.

Page 81
CVE 311 – Earthquake Engineering

3.6 Time History


The time history is the sequence of values of any time-varying quantity (such as a
ground motion measurement) measured at a set of fixed times. Also termed time
series.

To perform a time history analysis, a representative earthquake time history is


required for a structure being evaluated. Time history analysis is a step-by step
analysis of the dynamic response of a structure to a specified loading that may vary
with time. Time history analysis is used to determine the seismic response of a
structure under dynamic loading of representative earthquake (Wilkinson and Hiley,
2006)

The following is a snapshot taken from Philvocs:

Below is a sample of its details:


Page 82
CVE 311 – Earthquake Engineering

3.7 Requirements/Deliverables
Answer the following questions/problems:

1 What is the difference between earthquake intensity and magnitude?


2 What are the types of magnitude and explain?
3 What is attenuation?
4 What are the different attenuation relationships and explain?
5 The seismogram from an earthquake recorded by a seismograph located
exactly at 100 km from the earthquake epicenter exhibits a peak amplitude of
2 mm. Determine the magnitude of the earthquake.
6 Determine the peak ground displacement induced by an earthquake with a
magnitude of 8.5 at a distance of 100 km from the earthquake epicenter.

Page 83

You might also like