Introduction To Earthquake Engineering
Introduction To Earthquake Engineering
Engineering
Introduction to Earthquake
Engineering
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
This book contains information obtained from authentic and highly regarded sources. Reasonable efforts have
been made to publish reliable data and information, but the author and publisher cannot assume responsibility for
the v alidity of all materials or the consequences of their use. The authors and publishers have attempted to trace
the copyright holders of all material reproduced in this publication and apologize to copyright holders if permission
to publish in this form has not been obtained. If any copyright material has not been acknowledged please write and
let us know so we may rectify in any future reprint.
Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or
utilized in any form by any electronic, mechanical, or other means, now known or hereafter invented, including
photocopying, microfilming, and recording, or in any information storage or retrieval system, without written per-
mission from the publishers.
For permission to photocopy or use material electronically from this work, please access www.copyright.com (http://
www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA
01923, 978-750-8400. CCC is a not-for-profit organization that provides licenses and registration for a variety of users.
For organizations that have been granted a photocopy license by the CCC, a separate system of payment has been
arranged.
Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only
for identification and explanation without intent to infringe.
Chapter 1 Introduction...................................................................................................................1
1.1 Structural Effects of Earthquakes......................................................................2
1.1.1 Ground Failures..................................................................................... 2
1.1.2 Indirect Effects of Earthquakes............................................................5
1.1.3 Ground Shaking....................................................................................7
1.2 Types of Earthquakes.........................................................................................8
1.2.1 Man-Made Earthquakes........................................................................ 9
1.2.2 Volcanic Earthquakes.......................................................................... 11
1.2.3 Tectonic Earthquakes.......................................................................... 11
1.3 History of the Development of Mitigation Strategies for the Effects of
Seismic Hazards............................................................................................... 14
References................................................................................................................... 17
v
vi Contents
Hector Estrada
Luke S. Lee
ix
x Preface
MATLAB® is a registered trademark of The MathWorks, Inc. For product information, please contact:
xi
Authors
Hector Estrada, PhD, PE, is currently a professor of civil engineering at University of the
Pacific (Pacific), where he also served as chair of the Civil Engineering Department. Prior to join-
ing Pacific, Dr. Estrada was chair of the Department of Civil and Architectural Engineering at
Texas A&M University-Kingsville. Dr. Estrada has also had visiting research appointments at
NASA, Texas Tech University, and the U.S. Army Engineer Research and Development Center—
Construction Engineering Research Laboratory. His teaching interests include structural engineer-
ing and mechanics, the design of concrete, steel, and timber structures, structural dynamics, and
earthquake engineering. He has published on structural engineering and engineering education in
various peer-reviewed journals, conference proceedings, and presented research work at various
technical conferences. He has published a book on drafting and design of structural steel b uildings,
and several book chapters. He has served as a reviewer for a number of journals, conferences,
book publishers, and funding agencies. Dr. Estrada earned his BS (with honors), MS, and PhD all
in Civil Engineering from the University of Illinois at Urbana-Champaign.
Luke S. Lee, PhD, PE, is currently an associate professor of civil engineering at University
of the Pacific, where he teaches courses in solid mechanics, structural dynamics and health
monitoring, structural design, and engineering risk analysis. He has authored and co-authored
numerous journal articles, conference papers, and book chapters in composite materials, structural
health m onitoring, service life estimation, and engineering pedagogy. Dr. Lee earned his BS in
civil engineering from the University of California, Los Angeles, MS in civil engineering from
the University of California, Berkeley, and PhD in structural engineering from the University of
California, San Diego.
xiii
1 Introduction
“If a tree falls in a forest and no one is around to hear it, does it make a sound?” Since sound
is the sensation excited in the ear when air is set in motion, one could argue that the tree makes
no sound. Likewise, if an earthquake occurs and no one, or no infrastructure system, is there to
experience the resulting motion, we can conclude that the most severe earthquake would cause no
damage. Before humans appeared on earth, numerous earthquakes occurred, as evidenced by the
large shift in the location of the continents with no damage (this will be discussed in more detail
in Section 1.2.3). So we could argue that earthquakes do not cause damage; rather, it is infrastruc-
ture that experiences damage because of an earthquake. That is, the devastating effect of earth-
quakes should be attributed to an inferior built environment. Thus, when infrastructure is designed
to account for the effects of earthquakes, even the largest earthquakes cause little or no damage!
Nevertheless, because of the historical effects of earthquakes on human civilization, they are the
most terrifying of all natural disasters. The principle reason being that they can occur with little or
no notice and can devastate large areas.
This book is an introduction to the essentials of seismic design. The primary focus is on estab-
lishing the forces and deformations experienced by structural systems found in infrastructure. We
begin by providing a broad overview of earthquakes, their effects on structures, sources, and types
(Chapter 1), followed by an introduction to seismology in Chapter 2, the material for which was
selected based on seismology material covered in the California seismic portion of the Professional
Engineering (PE) exam. The book then covers the prerequisite structural dynamics material (based
on single-degree-of-freedom [SDOF] dynamic analyses) needed to develop the fundamental
relationships between structures and ground motion produced by earthquakes (Chapter 3).
The material being primarily focused on the development of parameters that are based on free and
forced, damped and un-damped vibrations, with the aim of developing parameters to characterize
structural system mass, stiffness, and damping, which in turn are used to calculate periods, frequen-
cies, and other relevant vibration properties.
We also explain the computational processes used to develop a critical concept (Chapter 4),
which is the establishment of seismic forces for structures at a particular location: the response
spectrum based on the Newmark–Hall method—both elastic and inelastic spectra (Chapter 5).
Next, we connect the analysis of multi-degree-of-freedom (MDOF) building systems (or multi-
story shear b uildings) to fundamental SDOF principles through the development of the general-
ized SDOF equations (Chapter 6). This generalized SDOF formulation can be combined with the
principles associated with SDOF analyses, including response spectrum analysis, to analyze a
multistory building as an SDOF system, including finding maximum shear and bending moment in
all members due to a seismic load. The process entails computing approximate participation factors
and using a design response spectrum to determine the story displacements, which are then used to
obtain lateral story seismic forces that are necessary to perform a complete analysis for the internal
loading of a structural system.
1
2 Introduction to Earthquake Engineering
Following the response spectrum analysis of MDOF systems subjected to seismic ground motion
in Chapter 7, we examine code-based analyses—both equivalent-static and dynamic approaches in
Chapter 8. The dynamic analysis will emphasize modal response spectrum analysis. The main goal
of the book is to elucidate code-based formulations such that a practicing engineer or an engineer-
ing student can apply these formulations, and explain how they were developed and describe the
limitations of these code-based procedures.
The book is intended as a tool for a one-semester course in introduction to earthquake engineer-
ing at the undergraduate or graduate levels, both in civil or architectural engineering.
The material presented in the book is concise and includes both essential structural dynamics
and seismic engineering principles, the former being prerequisite to fully grasping the funda-
mental concepts covered in the latter. Thus, students who have a limited background in structural
dynamics still have the opportunity to cover the essentials of earthquake engineering in a one-
semester course. This is particularly important in an era where information is readily available and
can be overwhelming for a new learner. For completeness, we also include computational tools that
can be used to solve the complex mathematics involved in the solution of even the simplest seismic
engineering problems.
At the beginning of each chapter, we present a list of learning outcomes in an effort to improve
the delivery of the content and to guide the learner in mastering the material presented. Overall, the
book adheres to the following general learning outcomes:
The first objective relates to the student’s ability to operate in the first two cognitive domains
of Bloom’s Taxonomy (namely, the knowledge of seismic engineering terms and comprehension of
the overall area of earthquake engineering). The second and third objectives primarily concentrate
on the next two cognitive domains (namely, the application of structural dynamics in earthquake
engineering and the analysis of systems subjected to earthquake loading).
(a) (b)
FIGURE 1.1 Surface rupture: (a) slow (creep) and (b) caused by the 1992 Landers earthquake in San
Bernardino County. (http://www.conservation.ca.gov/cgs/rghm/ap/Pages/main.aspx)
experienced over 16 inches (400 mm) of fault creep over a 40-year period as indicated by a graph of
fault slip displayed at the winery. Since the winery building straddles the surface of the fault, this
movement has been splitting it into two, which is clearly evident by the large crack on the concrete
floor. The San Andreas Fault, however, is capable of producing large earthquakes near this site,
such as the 1989 Loma Prieta earthquake, which caused widespread damage throughout the San
Francisco Bay Area, but did not break through to the surface. Figure 1.1b shows the surface rupture
resulting from the 1992 Landers earthquake, in San Bernardino County, California. The sudden slip
of the fault extended for nearly 50 miles with relative displacements from 1 inch to 20 feet.
Ground subsidence occurs as loose soils rearrange and settle into a denser state during vibra-
tions caused by earthquakes. In some cases, the compaction effect may amount to substantial
settlement of the ground surface. Liquefaction, which will be discussed later in this section,
can also trigger significant subsidence of unconsolidated sediment. For example, it is believed
that Big Lake and St. Francis Lake in Northeast Arkansas formed during the 1811 and 1812
New Madrid e arthquakes in areas where liquefaction resulted in significant lowering of the land
surface, causing it to flood. Uniform settlement over large tracks of land causes few problems
for infrastructure (except for long horizontal structures, such as roads and pipelines). However,
systems along the boundaries can experience differential settlement that can be extremely dam-
aging, as was experienced in Anchorage, Alaska along Fourth Avenue during the Good Friday
earthquake in 1964; see Figure 1.2.
Ground cracking is usually observed along the edges of ground subsidence as shown in Figure 1.2;
it may also be the result of slope failure or liquefaction, all of which cause the ground to lose its
support and sink, with the ground surface breaking up into fissures, scarps, horsts, and grabens
(Figure 1.2). The most damaging effect of ground subsidence is differential settlement, which can
severely disrupt the function of any infrastructure system near the vicinity of cracking locations,
particularly those with long foundations that straddle the cracks.
Soil liquefaction occurs when loose, saturated granular soils temporarily change from a solid to
a liquid state, losing their shear strength, which corresponds to a loss in effective stress between soil
particles. Loose saturated (or moderately saturated) sands and nonplastic silts are most susceptible
to this ground failure; however, in rare cases, gravel and clay can also experience liquefaction. In all
cases, poor drainage within the loose soil causes an increase in the pore water pressure as the soil
is compressed by the vibratory effect of seismic waves. As the load is transferred from soil to pore
water pressure, the effective stress between particles is temporarily reduced, or eliminated, causing
4 Introduction to Earthquake Engineering
Scarp
Fissure
Horst Graben
Normal
fault
a corresponding decrease in shear strength. In some cases, the pore water pressure increases rapidly
and a slurry (soil water mixture) forms that flows vertically to the surface, which can result in craters
and sand boils as shown in Figure 1.3a. Liquefaction is responsible for some of the most spectacular
failures caused by earthquakes as shown in Figure 1.3b. Following the 1964 Niigata, Japan earth-
quake, several apartment buildings experienced severe tilting and settlement. Fortuitously, the tilted
buildings did not suffer major structural damage, and thus minimal human injury.
Landslides caused by earthquakes are uncommon. Consequently, in order for a structure to
experience damage during the event, it must be located at the top or bottom of the soil mass that
slides down; for this reason, damage resulting from earthquake-induced landslides is rare. Sloped
land that is marginally stable under static conditions is most susceptible to sliding during the intense
shaking of strong earthquakes. For the most severe cases, debris (soil, boulders, and other mate-
rials) flow can move at avalanche speeds and can travel long distances depending on the slope
from which the landslide was formed. Furthermore, earthquake-induced landslides can be sudden
(a) (b)
FIGURE 1.3 (a) Sand boils that erupted following the 2011 Christchurch Earthquake and (b) Tilting and
settlement of buildings caused by liquefaction following the 1964 Niigata Earthquake. (Courtesy of USGS.)
Introduction 5
(a) (b)
FIGURE 1.4 (a) Partial view of Yungay City, Peru in 1966 and (b) after being buried by the 1970 Great
Peruvian earthquake-induced landslide. (Courtesy of EERC, University of California.)
and unpredictable, producing the total destruction of communities in the path of the debris flow.
Some devastating landslides have occurred around the world. One of the worst cases happened
in Peru during the 1970 Ancash (or Great Peruvian) earthquake (magnitude 7.9) off the Pacific
Ocean coast, which produced what is considered the world’s deadliest earthquake-induced land-
slide (20,000 fatalities). During the earthquake, the northern wall of Mount Huascaran, 130 km
(70 miles) from the epicenter, was loosened, mobilizing rock, ice, and snow into a massive landslide
that buried the towns of Ranrahirca and Yongay; see Figure 1.4.
BANGLADESH
INDIA
MYANMAR
(BURMA)
THAILAND
SOMALIA SRI
LANKA MALAYSIA
MALDIVES
KENYA
INDONESIA
TANZANIA
SEYCHELLES
Indian
ocean
MADAGASCAR
SOUTH
AFRICA
FIGURE 1.5 Countries affected by the 2004 Indian Ocean Tsunami. (From Wikipedia.)
as the fire consumed everything in its path. Traditional firefighting methods are often ineffec-
tive against earthquake-induced fires because most water mains that supply water hydrants break.
Earthquake-induced fires are started by ruptured combustible substance conduits (such as gas
mains) or destroyed combustible substance storage containers (such as oil tanks), and then ignited
by sparks from sources such as downed powerlines. Earthquake-induced fires are most common
in regions where timber construction is prevalent. One of the most devastating earthquake-induced
fires happened after the 1906 San Francisco earthquake. Over a four-day period, the fire burned
nearly 25,000 buildings on 508 city blocks (see Figure 1.6), which constituted about 80% of the
city at the time.
FIGURE 1.6 Extent of San Francisco fire following the 1906 Earthquake. (Courtesy of USGS.)
Introduction 7
1.
Earthquake size (magnitude): It can be measured objectively or subjectively—larger
earthquakes cause stronger shaking. A strong earthquake can cause ground shaking over
widespread areas, suddenly affecting large numbers of structures. Even relatively small
earthquakes can have a significant impact on large numbers of buildings. The 2014 South
Napa earthquake (magnitude 6), as shown in Figure 1.7, caused extensive damage to
buildings in the town of Napa, as well as to the many wineries in the region, with a total
estimated loss of approximately one billion US dollars. Fortuitously, the quake struck early
in the morning when businesses were closed and sidewalks were empty; only one fatality
was attributed to the quake.
2. Location (distance from the focus or epicenter): Generally, the closer to the epicenter, the
stronger the shaking. Structures near the epicenter of a strong earthquake can experience
extensive damage, in some cases partial or total collapse; see Figure 1.7. Many other
examples can be found online.
3. The subsurface materials beneath the structure: Soft soils amplify the shaking, while
rocks do not. This is the most insidious of the three factors because the site can be located
at a long distance from the epicenter and still experience extensive ground shaking due to
local soil conditions. Seismic waves travel through rock for most of their trip from the focus
to the surface; however, at many sites, the final part of the trip is through soil, the geologi-
cal characteristics of which have a major influence on the nature of ground shaking. Some
soils act as seismic wave filters, attenuating shaking at some frequencies while amplifying
it at others. The most peculiar example of distant seismic wave amplification is that of the
Valley of Mexico, where Mexico City is located. The origin of these waves (earthquake
focus) is approximately 350–400 km away, along the Pacific cost in Michoacán. Parts of
this valley are situated in a dry lake, Lake Texcoco, which was emptied to make way for
8 Introduction to Earthquake Engineering
38.5°
Saintsbury Winery
Sam Kee Laundry Building
km
37.5° 0 50
–123° –122°
Map version 18 processed 2014–08–25 04:21:30 AM UTC
Perceived shaking Not felt Weak Light Moderate Strong Very strong Severe Violent Extreme
Potential damage None None None Very light Light Moderate Mod/heavy Heavy Very heavy
Peak ACC (%g) <0.1 0.5 2.4 6.7 13 24 44 83 >156
Peak vel (cm/s) <0.07 0.4 1.9 5.8 11 22 43 83 >160
Instrumental
intensity I II–III IV V VI VII VIII IX X+
FIGURE 1.7 Intensity and extent of 2014 South Napa Earthquake. (Courtesy of USGS.)
the expansion of Mexico City. The near-surface geology of the city center consists of an
alluvial deposit, a thick layer of a very soft, high-water-content mixture of clay and sand
that was deposited there by streams before the lake was drained (so that the Aztecs could
build the island city of Tenochtitlan). As shown in Figure 1.8, rock shakes at the same
frequency as the seismic waves, which attenuate as they move further from the epicen-
ter. However, the unconsolidated alluvial soil has a dominant influence, and magnifies
the intensity of ground shaking (10-fold) near certain periods, resulting in near harmonic
motion with a period of 2 s. During the 1985 Mexico City earthquake (magnitude 8), peak
horizontal accelerations were first attenuated from 150 cm/s2 along the Pacific coast near
the epicenter to 35 cm/s2 at higher elevations with firm subsurface materials; these waves
were then amplified fivefold to 170 cm/s2 near the center of Mexico City. The shaking
intensity was further amplified by resonant vibration properties of certain buildings, those
with natural periods of about 2 s, which correspond to buildings between 10 and 20 sto-
ries, most of which experienced large displacements and extensive structural damage, or
collapse.
20°
–170
SCT 170 PGA = 35 cm/s2
16°
km
–170
Map version 1.1 processed Sat Nov 8, 2008 11:44:37 AM MST Campos Teacalco
Perceived shaking Not felt Weak Light Moderate Strong Very strong Severe Violent Extreme
Epicenter Mexico
Potential damage None None None Very light Light Moderate Moderate/heavy Heavy Very heavy city
Peak ACC (%g) <.17 .17–1.4 1.4–3.9 3.9–9.2 9.2–18 18–34 34–65 65–124 >124 ~400 km
Peak vel (cm/s) <0.1 0.1–1.1 1.1–3.4 3.4–8.1 8.1–16 16–31 31–60 60–116 >116
Instrumental
intensity I II–III IV V VI VII VIII IX X+
Seconds
0 10 20 30
11 Sep 2001
14 h 28 m 31 s
0
Ground velocity (µm/s)
–5
5 Station: PAL
17 Jan 2001 Component: E-W
12 h 34 m 22 s
FIGURE 1.9 Comparison of building collapse and natural earthquakes. (From Kim et al. 2001. Seismic
waves generated by aircraft impacts and building collapses at World Trade Center, New York City.
Eos Transactions American Geophysical Union 82(47): 565–571.)
Seismic activity near large artificial lakes has also been reported. It is believed that the weight
of the water and the extra water pressure in rocks can change the stresses in existing rock frac-
tures. The weight of the water increases the total stress by direct loading, while the extra pore
water pressure decreases the effective stress (pushes the fracture planes apart); this can lead to
sliding along fracture planes where tectonic strain is already present, causing an earthquake.
This is speculated to be most likely during periods when a reservoir is being filled or drained.
One case that continues to be attributed to artificial lake seismicity is the August 1, 1975 Orville,
California earthquake. Bolt (2004), however, presents a compelling argument against this sup-
position. His personal experience is that the connection between the earthquake and the fill-
ing of the lake is only circumstantial; however, he does not completely discount the possibility
that the reservoir might have in part triggered the earthquake. There are other earthquakes that
have been conclusively proven to have been triggered by artificial lakes; one such case is the
November 14, 1981 earthquake (magnitude 5.6) induced by Lake Nasser along the Nile River in
Egypt (Bolt, 2004).
Earthquakes caused by energy production processes have been recorded for over 50 years, though
recently the association of earthquakes and energy production has become contentious. This is
particularly controversial with respect to the hydraulic fracturing (fracking) process. This technique
for extracting oil and gas from low permeability rocks involves fracturing rocks by injecting high-
pressure fluid into the rock. The most compelling evidence of fracking-induced seismicity can be
found in Oklahoma and north Texas, regions where seismic activities were once nonexistent, but
now routinely experience small earthquakes; with the only variable being oil and gas production.
In fact, a 2012 report by the U.S. National Research Council concluded that some (not many) pro-
duction activities have induced perceptible seismicity. Geothermal power generation has also been
Introduction 11
conclusively correlated with earthquakes. The process entails injecting water at high pressures to
fracture 500°F solid rock, creating an artificial reservoir of superheated water, the steam of which
is then used to drive electrical turbines. A well-documented case of induced seismicity is from
Geysers Geothermal field in California.
1.2.3 Tectonic Earthquakes
As discussed previously, these are caused by a sudden dislocation of segments of the earth’s crust,
the structure of which is composed of plates (large and small) known as tectonic plates that float on
a liquid layer, the mantle. This arrangement resulted from the formation of planet Earth five billion
years ago, when hot gasses cooled into a semi-solid mass. It is estimated that after one to two billion
years of cooling, the crust solidified and cracked forming tectonic plates (different ones than those
that exist today).
From the beginning, the plates have been in constant motion forming and breaking up conti-
nents over time, including the formation of supercontinents that contained most of the landmass.
The latest supercontinent, Pangea, started separating approximately 200 million years ago, and
its parts have drifted apart to the current configuration of the earth’s surface. This process was
originally proposed by Alfred Wegener in the early 1900s. He noted several different pieces of
evidence to support his theory of the continental drift, including (see Figure 1.10):
1. How the current shape of some continents appear to fit together, particularly the east coast
of South America and the west coast of Africa.
2. The significant similarities between fossil records (both flora and fauna) found in several
continents that could only have occurred if the continents were attached.
3. The similarity in geology across several continents, including grooves carved by glaciers
and the sediments deposited by these glaciers.
Wagner’s theory implies that the earth’s crust (known as the lithosphere) consists of large slabs
of rigid rock, or plates (see Figure 1.11). The lithosphere can be divided into oceanic and continental
parts. The oceanic lithosphere is thinner and forms near mid-oceanic ridges where plates are drift-
ing apart. The thicker continental lithosphere plates consist of rocks that are less dense than oceanic
lithosphere rocks. All plates move at different speeds and different directions, spreading apart along
some of the plate boundaries (spreading or divergent zones), while colliding along other boundaries
(subduction or convergent zones). The spreading zones are continuously replaced with upwelling
lava, creating new seafloor or oceanic lithosphere. At the opposite side of the newly created oceanic
lithosphere, the old part of the plate plunges into the earth’s interior at the subduction zones, being
12 Introduction to Earthquake Engineering
SOUTH AMERICA
AUSTRALIA
ANTARTICA
FIGURE 1.10 Continental drift evidence; distribution of fossils across the southern continents of Pangea.
(Courtesy of USGS.)
absorbed into the mantel. At some locations along the mid-oceanic ridge, horizontal slip occurs
normal to the divergent zones as a new s urface is created or transformed; this type of boundary is
known as a transformed fault (see Figure 1.11).
Wagner originally proposed that the forces driving the continental drift were the result of g ravity,
and the same forces were responsible for ocean tides. This theory was rejected because these forces
are too small to drive the continents through the seafloor. It was later proposed that seafloor spread-
ing at the mid-oceanic ridges was responsible for moving the seafloor and continents over the
mantle, which is a softer rock immediately below. Although this theory was also rejected, it helped
establish the theory of plate tectonics, which with respect to the shape and size of the plates is
widely accepted today; see Figure 1.11. For a while, it was believed that the movement of the plates
was driven by convection currents in the mantle; this transfers heat from deep within the earth (the
core) to the surface. The temperature difference drives molten lava currents (similar to how heat
from the sun drives wind currents on the earth’s surface). The currents are produced as heated rock
near the core (which has a lower density than that near the bottom of the plates) rises toward the
surface as molted lava; over the course of the trip, lava cools and becomes denser, some of which
after reaching the surface plunges back toward the core.
This theory, however, does not fully explain all phenomena associated with lithospheric plate
movement. The current view is that plate movement is driven by additional forces, known as ridge
push and slab pull, which also derive their energy from the dissipation of heat from the earth’s core
through convection currents (this theory is still being debated). These forces are generated at the
mid-oceanic ridge (divergent zones) and at the subduction zones. At the divergent zones, lava cools
at the surface creating new oceanic ridge; as this new material moves away from the divergent
Introduction 13
Plate
Plate
Asthenosphere
North American
Eurasian plate
plate Eurasian
plate
Juan de Fuca
plate
Caribbean
Filipino plate Arabian
plate plate Indian
Cocos
plate plate
EQUATOR
Nazca African
Pacific
plate plate
Australian plate South American
plate plate
Australian
plate
Scotia plate
Antarctic
plate
FIGURE 1.11 Tectonic plates and types of plate boundaries. (Courtesy of USGS.)
zones, it continues to cool, becoming denser. At the other end of the plate, the oceanic lithosphere
becomes even denser with age as it continues to cool and thicken. This causes the lithospheric plate
to tilt toward the subduction zones, like an unbalanced boat in water. The mid-oceanic ridge then
rises above the rest of the seafloor, and as the new material rises to the top of the ridge, gravity pulls
it down; this is called ridge push. Also, the greater density of old oceanic lithosphere at the subduc-
tion zones, relative to the underlying asthenosphere, allows gravity to pull these rocks (known as
slabs) down into the mantle. This is believed to be the source of the driving force for the majority
of plate movement.
Although the precise force mechanisms causing movement of the lithospheric plates con-
tinue to be debated, there is substantial agreement that heat from the earth’s core serves as the
primary source of energy for these forces. Also, the general outlines of the major plates have
almost conclusively been established. In fact, this outline is clearly illustrated when the loca-
tion of the epicenters of recent earthquakes is plotted on a global map as shown in Figure 1.12.
At the boundaries of the plates, rocks fracture, usually at many locations, creating a web of
smaller plates with edges that rub and push relative to each other; these edges are called faults.
On average, these faults have the potential to displace approximately 2 inches per year. When
the rubbing and pushing is prevented, elastic energy accumulates along the edges of the plates.
When this energy is released with a s udden movement (slip), it causes brief strong ground vibra-
tions. The specific location (generally a volume of rock) where the movement or energy release
occurs is known as the focus, or hypocenter. The point on the earth’s surface directly above
the hypocenter is called the epicenter. Usually, the vibrations cause the rocks near the focus to
become unstable; and as these rocks settle into a new equilibrium state they cause aftershocks.
The discipline that studies seismic activity is known as seismology, and will be discussed further
in Chapter 2.
14 Introduction to Earthquake Engineering
African plate
While humans did not fully understand the process that causes earthquakes until the middle
of the last century, many civilizations developed construction practices to reduce the s tructural dam-
age caused by ground shaking. From the study of prehistoric buildings in seismic-prone regions, it
appears that two approaches to seismic-resistant construction were followed: increase the strength
or change the stiffness of the lateral force-resisting systems. This is particularly evident for the con-
struction of important structures, which received special seismic-resistance detailing, while most
other buildings did not.
In many parts of the world, buildings were built stronger after they collapsed in an earthquake,
while other parts of the world addressed weaknesses in the construction by specifying different
construction detailing. For example, the Incas in Peru were master builders that created ingenious
earthquake-resistant structures. The hilltop city of Machu Picchu (classified as a Civil Engineering
Introduction 15
International Historic Monument by the American Society of Civil Engineers [ASCE]) has numer-
ous structures constructed of interlocking, mortar-free stonework with perfectly mated joints (these
are known as dry-stone walls or ashlar masonry); see Figure 1.13. Clearly, this type of construc-
tion was deliberate in order to prevent sliding of the stone blocks during earthquakes. The intricate
process was far more complex than the perfectly regular array of stonework used in other parts
of the world, indicating that the extra effort was purposefully made for structural reasons and not
for aesthetics, as some have suggested. In fact, there are buildings in Machu Picchu with masonry
laid in parallel courses (see Figure 1.13); the most important structures, such as temples and the
king’s quarters, were built to withstand large earthquakes using masonry laid in nonparallel courses
(see Figure 1.13).
Ashlar masonry is more seismic resistant than using mortar because higher friction forces
develop between mating surfaces. This along with the interlocking masonry allows the stones to
move slightly and resettle without excessive movement of the walls. This type of construction allows
for passive energy dissipation, a form of passive structural control. Inca construction was based on
designs that follow prescriptive specifications. This type of design philosophy is based on experi-
ence with the construction of proven designs, and is still widely used for design today. More rational
construction procedures based on engineering principles (engineered construction) have only been
available for the past century.
Engineered seismic-resistant design was developed in America, Asia, and Europe; not neces-
sarily independently, but there is no consensus on the order. It is believed that the first rational
attempts to understand the effects of earthquakes on structures were made in Europe following
the devastating 1755 Lisbon earthquake (over 50,000 casualties). After this event, the regional
government enacted building regulations that required seismic-resistant detailing to be followed
in the reconstruction of the city. A few years later, scientists thoroughly documented the effects of
the 1783 earthquake in the Calabria region of southern Italy (nearly 50,000 casualties). Although
the results from these studies did not lead to substantive design guidelines, they did inspire future
scientists to study the effects of earthquakes on structures, including a number of British and Italian
engineers who went on to develop the foundation for seismology.
After the 1881 Nobi earthquake (7000 casualties), the Japanese government formed the first-ever
earthquake investigation committee, and directed the committee members to research earthquake-
resistant structural design. It is believed that this committee conceived the idea of characterizing
the effects of earthquakes on structures by applying a lateral force equal to a fraction of the total
weight of the building, which is based on Newton’s second law (force equals mass times accelera-
tion). Following the 1908 Messina-Reggio earthquake (58,000 casualties), the Italian government
also appointed a committee to develop practical recommendations for the seismic-resistant design
of buildings. This committee also proposed characterizing the effects of earthquakes on structures
by applying lateral forces: first, they proposed a force equal to 1/12 of the building weight. Three
16 Introduction to Earthquake Engineering
years later, this specification was modified as follows: the first story must resist a force equal to 1/12
of the building weight above, while the second and third floors must resist a lateral force of 1/8 of
the building weight above. This specification limited the height of buildings to three stories. These
two committees provided a first iteration to the equivalent static procedure that is widely used in
seismic-resistant design codes around the world today.
In America, the first attempt to characterize seismic effects on buildings also followed a major
seismic event, the 1906 San Francisco earthquake. Resistance to earthquake forces was prescribed
in code provisions indirectly by requiring buildings to resist a wind lateral load of 30 lb/ft2, which
was intended to account for both wind and seismic effects. Following another devastating earth-
quake in California (Long Beach in 1933), the Division of Architecture was given the responsibility
for public-school building safety. Their proposed regulations required that public-school buildings
be designed to resist a lateral force equal to a fraction of the dead load and a portion of the design
live load, entailing 0.1 for masonry buildings without frames and 0.02–0.05 for other types of
buildings. This appears to be the first explicit design specification incorporating the importance of
construction regarding the type of lateral force-resisting system. This also marked a very impor-
tant point in code enforcement; the California Legislature passed the Field Act and the Riley Act.
The Field Act made the seismic-resistant design and construction of public-school buildings manda-
tory, and the Riley Act required that most buildings in the state be designed to resist a lateral load
equal to 2% of the design loads dead plus live loads.
The 1943 City of Los Angeles building code is believed to be the first to recognize that lateral
forces depend on building mass and height, an indirect recognition of building flexibility on these
forces. The lateral forces at different stories of a building were determined by the product of a grav-
ity load (equal to dead load plus half of the live load) and a lateral force coefficient that varied with
the number of stories above the story being analyzed. The 1947 San Francisco building code also
recognized the influence of building flexibility on the magnitude of seismic forces, as well as the
influence of site soil conditions on these forces.
Although many city building codes have been in existence around the world for decades (even
millennia, such as the Hammurabi Code from Babylon, which did not include specifications, rather
promoted good construction by prescribing penalties for failures), regional model codes were not
developed until the introduction of the Uniform Building Code (UBC) in 1927. The second edition
of the UBC published in 1930 incorporated seismic provisions and suggested these be incorporated
by cities in areas prone to earthquakes. The UBC seismic provisions were based primarily on two
observations of building performance during earthquakes: first, most damage was caused by lat-
eral shaking, and second, structural damage was more extensive for buildings on soft soil deposits
(likely the first code to recognize soil effects). The code specified a lateral force equal to 10% of
the dead plus design live loads, reducible to 3% for buildings on firm soil. In the ensuing editions
of the UBC, maps of the United States depicting zones of different levels of seismic risk were
published. These maps recognized the hazard of large magnitude earthquakes for different areas
of the country.
In 1948, a joint committee of the Structural Engineers Association of Northern California and
the San Francisco Section of ASCE proposed model lateral-force provisions for California building
codes. A major advance was the introduction of the earthquake-response spectrum concept, which
was first developed by Biot in 1943 and later proposed as a design tool by Housner. This theory
combines ground motion and the dynamic properties of the structure (namely the period). This led
to the building period becoming an explicit factor in the determination of seismic design forces.
The response spectrum work was made possible by recordings of strong-motion earthquake accel-
erations using recently deployed accelerographs by the US Coast and Geodetic Survey. In 1952,
a second joint committee of ASCE and now the Structural Engineers Association of California
(SEAOC) developed a comprehensive guide, the Recommended Lateral Force Requirements and
Commentary (the so-called SEAOC Blue Book), first published in 1959. This explicitly related lat-
eral forces to building periods based on the design response spectrum.
Introduction 17
Structural dynamic concepts were further developed in the 1960s and incorporated in seismic
codes after the 1971 San Fernando earthquake. In particular, structural dynamic specifications
explicitly incorporating the response spectrum theory were required for the design of hospitals
and other critical facilities. Since the incorporation of structural dynamics in code specifications,
significant advances have been made in the development of analytical procedures; much of this has
been in parallel with the development of computers because solutions based on structural dynamics
are computationally intensive. Even with the great advances in the development of mitigation strate-
gies for ground shaking, structures designed using modern seismic-resistant design principles have
been observed to experience much larger forces during major earthquakes than those prescribed by
design codes, leading to major damage, even collapse. These observations have led researchers to
conclude that structures subjected to larger imposed loadings from earthquakes can survive intense
shaking with damage; however, will not collapse as long as lateral force-resisting systems are tough
(i.e., able to absorb the seismic energy) and include a continuous path for lateral load to be trans-
ferred to the ground. Also, it is important for engineers to understand that the structural engineering
principles used for the analysis of structures subjected to gravity static loads do not result in the
greatest dynamic forces a structure may experience during an earthquake. These dynamic forces
are only equivalent to code-specified lateral forces in that a structure designed to resist these forces
has the capability of deforming without overstressing from load reversals, and provide adequate
member ductility, as well as provide connections with sufficient strength and resiliency to accom-
plish the following performance goals:
These performance criteria have generally remained unchanged since the 1970s, but have
recently taken on a broader approach in order to allow the real estate owner to decide the desired
level of structural performance and safety; both in terms of protecting occupants and capital invest-
ment. This new philosophy is known as performance-based design and is awaiting to appear in stan-
dards such as the 2010 ANSI/ASCE “Minimum Design Loads for Buildings and Other Structures”
standard; see Section 1.3.1.3 in this document.
REFERENCES
Bolt, B. A., Earthquakes, 5th edition, W.H. Freeman and Co., New York, 2004.
Wald, D., Quitoriano, V., Heaton, T., and Kanamori, H., Relationships between peak ground accelera-
tion, peak ground velocity and modified mercalli intensity in California, Earthquake Spectra, 15(3),
557–564, 1999.
2 Engineering Seismology
Overview
After reading this chapter, you will be able to:
Seismology is the scientific discipline that studies seismic activity from various sources
(tectonic, volcanic, etc.), in particular, the propagation of seismic waves through the interior of the
earth and their environmental effects, such as ground shaking and tsunamis. This discipline grew out
of the need to understand how earthquakes occur, and to try and predict their occurrence. Engineering
seismology is the applied branch of seismology that focuses on characterizing seismic hazards at a site
or region in order to evaluate risks to various vulnerable infrastructure systems. This is accomplished
by studying the seismic history of a region (including the frequency and intensity of earthquakes)
to establish characteristics (such as the intensity of shaking) and frequency of occurrence of future
earthquakes. In this chapter, we briefly cover the most important concepts of engineering seismology.
19
20 Introduction to Earthquake Engineering
Epicentral distance
North Epicenter
Strike angle
Observing station
Dip angle
Fault line
Hypocenter (focus)
16 km (10 miles). Most moderate-to-large shallow earthquakes are preceded by smaller quakes,
called foreshocks, and followed by smaller quakes called aftershocks. Foreshocks are precursors of
the impending fault rupture, while aftershocks result from adjustments to the stress imbalance in
the rocks produced during the rupture.
The fault slip is used to classify faults depending on the direction of the movement of rocks on one
side relative to the other side; vertical movement is known as dip-slip, while horizontal movement
is known as strike-slip. These terms correspond to the definition of the angles used to describe the
direction of fault movement. As discussed in Section 1.2.3, faults occur along the edges of tectonic
plates (or their interior, which is much less common). Dip-slip faults are further subdivided based on
the relative vertical movement of the tectonic plates: normal faults occur at plate boundaries where
tectonic plates spread apart at divergent zones causing the hanging wall to move down relative to the
footwall, while reverse faults occur at plate boundaries where tectonic plates collide at convergent
zones causing the hanging wall to move up relative to the footwall; see Figure 2.2. Thrust faults
are reverse faults with a small dip angle. Faults along oceanic ridges are normal, resulting in the
lengthening of the crust, as shown previously in Figure 1.11, whereas those along subduction zones
are reverse, resulting in the shortening of the crust. When these faults break through to the surface,
surface rupture, they produce an exposed steep slope known as the fault’s scarp; see Figure 2.2.
At transform zones, tectonic plates move horizontal relative to each other; the faults associated
with this motion are strike-slip, which can be further subdivided into right-lateral or left-lateral
faults. Figure 2.3 depicts a left-lateral fault, which can be described by an observer standing in front
of the fault line as watching the land on the other side moving to the left. Similarly, an observer
facing a right-lateral fault line would see the land on the other side moving to the right. The San
Andres fault discussed in Chapter 1 is a type of right-lateral fault. This is demonstrated by the fault
extending from left to right through the drainage culvert on the south side of the DeRose Winery
building in the Cienega Valley; see Figure 2.3. Therefore, when these faults break through to the
surface, any pipes, streams, fences, etc. that straddle the surface rupture are offset. It should be
noted that dip-slip and strike-slip faults can occur simultaneously, in which case the fault is clas-
sified as an oblique-slip. Also, fault rupture does not reach the surface in the case of many earth-
quakes, particularly for intermediate and deep cases.
Scarp
Hanging
Footwall wall
Hanging Footwall
wall
Normal fault Reverse fault
Right-lateral fault
Left-lateral fault
2.1.2 Seismic Waves
Seismic waves radiate from the focus and travel in every direction, as shown in Figure 2.5. The portion
of the energy released from a fault rupture as shaking first travels through the interior of the earth
as body waves, following the shortest path to the ground surface where they are transformed into
surface waves (see Figure 2.5). Thus, body waves arrive at an observation station (or site) before any
surface waves. An understanding of these different waves is necessary to establish the epicenter and
to characterize the size of the earthquake (earthquake magnitude and ground acceleration).
Surface
waves
S Body waves
Earthquake
P
Vibration
direction
Body waves naturally divide into two different types, P-waves (primary or pressure) and
S-waves (secondary or shear), which are slower than the P-waves; see Figure 2.5. The stiffness of
the material through which the waves travel controls their speeds. P-waves are longitudinal waves
that move through earth’s interior in successive compression and rarefaction (i.e., P-waves move
material particles back and forth parallel to their direction). Their average speed through solids
is approximately 6 km/s and slows down when traveling through liquid, for example, speeds in
water are only about 2 km/s. Also, given that liquid medium (such as water) has no shear stiffness,
S-waves can only travel through solids, at about 3 km/s. Because of the slower speeds of S-waves,
they travel at a lower frequency, but larger amplitude, which makes them more destructive than
P-waves. This speed differential between P- and S-waves can be recorded using a seismograph as a
time lapse, Δtp−s (see Figure 2.6), and allows the seismologist to predict an impending earthquake
at a site, albeit with a very short warning. This was the basis for an earthquake early warning
system implemented in Mexico City a few years after the 1985 earthquake. Because of the large
distances (approximately 400 km; see Figure 1.8) from the epicenter to the city, it is estimated that
the residents of the city have 60–90 s (the time it takes between the arrival of the P-waves and the
intense shaking from S-waves) to prepare for the impending earthquake.
When body waves reach the ground surface, they are classified as surface waves. These have
lower frequency than their parent body waves and are obviously responsible for the damage associ-
ated with earthquakes. The intensity of these waves diminishes approximately exponentially as the
focal depth increases. These waves are also divided into two types: Rayleigh and Love. Rayleigh
waves combine effects of both P- and S-waves on the earth’s surface and move disturbed material
vertically and horizontally (in the direction of wave propagation) simultaneously, similar to ocean
waves. These waves have low frequencies and large amplitudes, and travel at average speeds of
Time
P-wave arrival
S-wave arrival
3 km/s. Love waves move the ground side to side on the surface plane, perpendicular to the direc-
tion of the wave propagation, similar in character and speed to S-waves. The shear motion of these
waves is primarily responsible for the ground shaking experienced during earthquakes.
As shown in Figure 2.6, seismic waves can be recorded using a seismograph, which can be used
to decipher the four different types of waves since each travels at different speeds, with P-waves
being the fastest, followed by S-waves, then Love waves, and finally Rayleigh waves. Along with
the direction of shaking on the ground surface (vertical and horizontal), the different speeds relative
to one another can be used to separate these waves from one another in a predictable pattern; this
is particularly clear for P- and S-waves in Figure 2.6. For surface waves, the vertical and horizontal
(north–south and east–west) recordings are needed to decipher the two surface waves. The different
arrival times of these waves can be used to estimate the epicentral distance (and the hypocentral
distance—the distance from the observation station to the focus). With recordings from three differ-
ent observation stations, an estimate of the geographic location of the epicenter can be triangulated.
That is, the epicenter is located where three circles drawn on a map intersect, each circle of radius
equal to the epicentral distance, and centered at their corresponding observation station location.
2.2.1 Earthquake Intensity
The intensity of an earthquake is a subjective, nonempirical approach for estimating the size of an
earthquake based on the subjective assessment of human observations of the effects of earthquake
shaking on buildings (amount of damage sustained by structures and land surface) and on people.
Different buildings (based on type and construction material) respond differently to shaking; that
is, the performance of poorly designed and constructed buildings is more sensitive to shaking than
1933
Observation (subjective) Measured motion (objective)
Timeline
well-designed and/or well-constructed structures. Also, people have varying perceptions of shaking,
that is, psychologically some people are more sensitive to shaking than others. This subjective measure
of earthquake intensity was developed before any instruments were capable of recording earthquake
shaking, and thus its value has diminished since the development of instrumental earthquake size
measurements. However, the intensity methodology is still useful for assessing historic events from
descriptions of damage to buildings and land, particularly given criteria that account for construction
quality, g eologic conditions, and potential discrepancies between observers of the same earthquake.
One of the earliest criteria (or scales) was a 10-degree scale proposed in 1883 by two scientists
for whom the scale was named the Rossi–Forel (RF) scale. This is an arbitrary scale that character-
izes shaking numerically in Roman numerals from 1 (RF-I) for barely noticeable to 10 (RF-X) for
an extremely intense earthquake; see Figure 2.7. This scale was not widely accepted; however, it
is still occasionally used in some parts of the world. In 1902, Giuseppe Mercalli proposed a modi-
fied version of the RF scale that found wide consensus amongst seismologists. The main issue with
the RF scale was that the 10th degree was too broad (see Figure 2.7). The new scale still had 10
degrees, but intensities were distributed more evenly, with more detailed and explicit definitions
of each degree. This new scale was later modified by a number of seismologists, including Richter
who modified it to better account for contemporary construction in California, and was published in
1931 as the modified Mercalli (MM) scale (Richter, 1958). This is still widely used today, and it has
12 points, ranging from 1 (MM-I) for imperceptible shaking to 12 (MM-XII) for total destruction;
see Table 2.1. It is important to note that even this scale has many limitations because it is based
TABLE 2.1
Modified Mercalli Intensity Scale, and Comparison to Magnitude and PGA
Degree Intensity Description Magnitude PGA ( g)
I Not felt, except by some in rare circumstances. <3.0 <0.03
II Felt only by a few persons at rest on upper floors of buildings. 3.0–3.9
III Felt indoors, but many people do not recognize it as an earthquake.
IV During the day felt indoors by many, outdoors by few, while at night some 4.0–4.9
awakened.
Dishes, windows, doors are disturbed and walls make creaking sounds.
V Felt by nearly everyone during the day and many awakened at night. Some dishes, 0.03–0.08
windows, etc. are broken; cracked plaster in a few places; unstable objects
overturned.
VI Felt by all, many frightened and run outdoors. Some heavy furniture moved; a few 5.0–5.9 0.08–0.15
instances of fallen plaster and damaged chimneys. Damage slight.
VII Everybody runs outdoors. Damage negligible in buildings of good design and 0.15–0.25
construction; slight to moderate in well-built ordinary structures; considerable in
poorly built or badly designed structures; some chimneys damaged.
VIII Damage slight in specially designed structures; considerable in ordinary buildings 6.0–6.9 0.25–0.45
(some with partial collapse); great in poorly built structures. Chimneys, factory
stack, columns, monuments, walls toppled. Heavy furniture overturned.
IX Damage considerable in specially designed structures; well-designed frame 7.0–7.9 0.45–0.6
structures thrown out of plumb; great damage for others, including partial
collapse. Buildings shifted off foundations. Ground cracked conspicuously.
Underground pipes broken.
X Some well-built wooden structures destroyed; most masonry and frame structures 0.6–0.8
destroyed. Ground badly cracked.
XI Few masonry structures remain standing. Bridges destroyed. Broad fissures in >8.0 0.8–0.9
ground. Underground pipelines completely out of service.
XII Damage total. Waves seen on ground surface. Objects thrown into the air. >0.90
Engineering Seismology Overview 25
2.2.2 Earthquake Magnitude
Whereas the intensity of a given earthquake varies from one observation point to another, earth-
quakes can be associated with a single value of magnitude. The consensus measure of magnitude is
based on the Richter scale, which quantifies the size of an earthquake with an index of the amount
of energy released, and while this approach is an improvement as compared to the intensity scales,
the Richter scale does not accurately account for all factors that contribute to the actual size of an
earthquake. This scale, however, does appropriately measure the relative strength of an earthquake
and remains an important parameter in earthquake hazard analysis. In addition to being used for
earthquake hazard analysis by seismologists and engineers, it is the preferred scale used to inform
the public of the size of an earthquake.
Because it was originally developed to quantify the strength of Southern California earthquakes,
the Richter scale is also known as the local magnitude scale, M L . Richter defined M L using the
base-10 logarithm of the peak trace amplitude (in micrometers, µm) of a standard Wood–Anderson
seismograph (which has a magnification factor of 2800, a natural period of 0.8 s, and damping of
80%), located on firm ground at a distance of 100 km from the epicenter. The following relationship
gives the local magnitude:
A
M L = log10
Ao
where:
A is the peak amplitude in µm, measured from a seismogram
Ao is the peak amplitude of a zero-magnitude earthquake in µm, which is used to adjust the
variation of ground motion amplitude for epicentral distances other than 100 km
This equation is typically not used directly to determine the magnitude of an earthquake; instead,
a correction nomogram provided by Richter is used (see Bolt 2004). The nomogram was convenient
when digital calculators were not available. The equation used to develop the nomogram is
where:
A is now given in millimeters (mm)
Δtp−s is the time between the arrival of P- and S-waves in seconds (see Figure 2.6); this
indirectly measures the epicentral distance
26 Introduction to Earthquake Engineering
The Richter scale has a practical range from 0 to 9.0 (theoretically, the scale has no upper or
lower limits). Also, the base-10 logarithmic scale indicates that each unit increase in M L corresponds
to a 10-fold increase of the earthquake wave amplitude. For example, a 7 M L earthquake is 100
times stronger than a 5 M L event (10 × 10 = 100).
EXAMPLE 2.1
Estimate the local magnitude of a southern California earthquake recorded in two perpendicular
directions at several stations using standard Wood–Anderson seismographs. The trace amplitudes
and epicentral distances are as follows (Richter, 1958):
Solution
Equation 2.1 can be rewritten in a slightly different form to include the epicentral distance rather
than Δt p–s, ML = log10 A + 2.56 log Δ − 5.12 (this type of equation is dependent on the seismologi-
cal station and still yields significant scatter of the data). This equation can be used to determine
ML for the two perpendicular directions, which can then be averaged to obtain a sufficiently
accurate result without having to combine the results vectorially. The calculations for Station 1 in
the north–south and east–west directions are as follows:
MLN−S = log10 (8.4 ×103 μm) + 2.56 log10 (114 km) − 5.12 = 4.07
MLE−W = log10 (6 ×103 μm) + 2.56 log10 (114 km) − 5.12 = 4.92
4.07 + 3.92
ML1 = = 4 .0
2
All other stations are treated similarly and the results are summarized as follows:
Local Magnitude
Therefore, the magnitude of this particular earthquake can be determined as the average of the
averages of each station:
4 . 0 + 4 . 6 + 4 . 3 + 4 .9
ML = = 4 .4
4
Engineering Seismology Overview 27
The local energy associated with earthquakes during fault fracture growth is primarily trans-
formed into heat, with only 1%–10% being released as seismic waves. The relationship between the
fraction of energy radiated as waves and the local magnitude, M L, is given as follows:
log10 E = 11.8 + 1.5ML (2.2)
where:
E is in ergs, which is a relatively small unit, 1 ft lb = 1.356 × 107 ergs
In Equation 2.2, the logarithmic scale coupled with the other terms results in a 32-fold increase
in energy radiated for each unit increase in M L . For example, a 7 M L earthquake radiates 1000 times
more energy than a 5 M L event (32 × 32 = 1024). Since this relationship was calibrated using vari-
ous types of seismic waves, it is also applicable to other magnitude scales, including those discussed
in the following paragraphs.
The local magnitude scale does not accurately account for deep focal depths (>45 km) or long
epicentral distances (>600 km) because seismic waves attenuate with distance (whether the distance
is depth or length). The wave train recorded on a seismograph for a deep earthquake is very differ-
ent than for a shallow one, leading to two different values of M L even for events that release equal
amounts of total energy. Consequently, more accurate measures of magnitude had to be developed
in order to improve the uniform coverage of earthquake size. Richter working with his colleague
Gutenberg addressed the shortcomings of the local magnitude scale by developing the body-wave
scale, mb, to handle deep-focus earthquakes and the surface-wave scale, Ms, to handle distant earth-
quakes. The three scales are related by the following empirical relationships:
and
Unfortunately, even these scales do not provide an accurate estimate for the largest magnitude
(>8M L ), most devastating earthquakes. The limitations of the scales are purported to be the result
of ground-shaking characteristics (such as the natural period) not increasing proportionally with
increasing amount of the total energy radiated by very large earthquakes. M L, mb, and Ms are valid
for natural periods of up to 0.8 s (for a Wood−Anderson seismograph), 1.0 s, and 20 s, respectively.
Because of these ground-shaking limitations, earthquakes larger than certain sizes will have a con-
stant magnitude. This is known as magnitude saturation. For instance, M L begins to saturate at
about 6.5, mb at about 7.0, and Ms at about 8.0.
The strength of an earthquake can be more accurately measured using the moment magnitude
scale, MW, which accurately measures a wide range of earthquake sizes and is applicable globally.
This scale is a function of the total moment release by an earthquake. The moment is a measure of
the total energy released. The concept of moment is adopted from mechanics and is defined as the
product of the fault displacement and the force causing the displacement. A simple derivation of the
moment, M0, is presented in Villaverde and is based on the size of the fault rupture, the slip amount,
and the stiffness of the fractured rocks. That is,
M 0 = GAf Ds (2.5)
where:
M0 is in dyne cm, which is a relatively small unit, 1 dyne cm = 1 × 10−7 N m
G is the shear modulus of the rocks included in the fault in dyne/cm2, which ranges from
3.2 × 1011 dyne/cm2 in the crust to 7.5 × 1011 dyne/cm2 in the mantle
28 Introduction to Earthquake Engineering
All of which can be estimated relatively accurately. Alternatively, the moment can be directly
estimated from the amplitudes of long-period waves at large distances, with corrections for attenu-
ation and directional effects. Although moment is an effective way to establish the size of an earth-
quake, it is customarily convenient to convert it into a magnitude quantity so that it can be compared
to M L, mb, and Ms. This can be accomplished by relating the seismic moment M0 to the radiated
energy E. That is,
∆σ
E= M0 (2.6)
2G
where:
Δσ is the static stress drop in the earthquake, which ranges from 30 to 60 bars (1 bar = 100 kPa)
G is the shear modulus of the medium near the fault, the same as used in Equation 2.5
The moment magnitude, Mw, is then derived by substituting Equation 2.6 into Equation 2.2 and
replacing M L with Mw since the two scales give similar values for earthquakes of magnitude ranges
from 3 to 5. For average values of Δσ and G (Δσ/G ≅ 10 –4), the relationship between M0 and Mw is
then given as
2
Mw = log10 M 0 − 10.7 (2.7)
3
where:
M0 is in dyne cm
The resulting Mw does not saturate at large magnitude values. Mw is also known as the Kanamori
wave energy, after the scientist who developed this relationship (Villaverde, 2009).
EXAMPLE 2.2
Estimate the seismic moment and moment magnitude of the January 12, 2010 Haiti earthquake.
It is estimated that the blind thrust fault (the slip plane ends before reaching the earth’s surface)
caused an average strike-slip displacement of 2 m over an area equal to 30 km long by 15 km
deep (Eberhard et al. 2010). Assume that the rock along the fault has an average shear rigidity of
3.2 × 1011 dyne/cm2.
Solution
1. Determine the fault’s rupture area A f and fault slip, Ds, in consistent units.
M0 = GAf Ds
= (3.2×1011 dyne /cm2 )( 4.5 ×1012 cm2 )( 200 cm)
= 2.88 ×10 26 dyne cm
Engineering Seismology Overview 29
Mw = 2/ 3log10M0 – 10.7
= 2/ 3log10 (2.88 ×10 26 dyne cm) – 10.7
= 6.94
It is worth noting that a magnitude 7 Mw has been widely reported in the media.
There are other magnitude scales in use that are not covered in this section; thus, it is important
to specify the type of magnitude scale being used. In this book, M L is used unless otherwise noted.
2.3.1 Earthquake Characteristics
Although the primary factor controlling an earthquake’s effect on structural behavior at a particular
site is the PGA, the largest PGA earthquakes do not always cause the most damage from shaking.
Shaking is rather complex and it is difficult to characterize its full effect using a single parameter.
Strong shaking also significantly depends on duration and frequency content, and to a lesser extent
on factors such as length of fault rupture, focal depth, orientation of the fault, speed of rupture, and
whether or not fault rupture reaches the ground surface.
As discussed earlier, the size of an earthquake can be measured using various parameters
(intensity, magnitude, etc.), the most important of which is PGA when it comes to seismic-resistant
design. This is obtained from the acceleration time history recorded with an accelerometer (which
is a recording of the variation of acceleration amplitude with time) and is typically given as a
fraction of the acceleration due to gravity. PGA is generally largest near the epicenter and tends to
decrease away from it because of attenuation. However, PGA can increase significantly depending
on a number of factors, especially the geology of the path of the waves to a specific site and local
soil conditions (to be discussed in the next section). Ground acceleration is generally resolved into
a vertical and two horizontal components. The vertical component is approximately one-third of
either of the horizontal components; thus, the majority of the damage is caused by the horizontal
components of acceleration. Nonetheless, the vertical component of acceleration is also included in
modern seismic design codes, as discussed in Chapter 8.
As mentioned earlier, large PGA earthquakes do not always create the most damage. In fact, a
large PGA earthquake would produce little damage to some structures if it only occurs for a short
time, whereas a relatively small PGA earthquake that continues for several seconds can be devas-
tating to some structures. This can be exemplified using the first two earthquake records in Figure
2.8; whereas the PGA for the 1940 El Centro earthquake (PGA = 0.32g, M = 7.0) is smaller than
30 Introduction to Earthquake Engineering
that from the 1966 Parkfield earthquake (PGA = 0.5g, M = 6.2), the resulting damage from the El
Centro earthquake was much more extensive than that associated with the Parkfield earthquake.
This can largely be attributed to the much longer strong-motion duration of the El Centro earth-
quake. The duration of strong motion (or shaking) is important because the longer the strong motion,
the more energy is imparted to the structure; and because of the limited capacity of a structure to
absorb energy, there is a greater potential for inelastic response for longer strong shaking. Since
inelastic damage accumulates from load reversals, an earthquake with a large number of load rever-
sals would result in more damage compared to an earthquake with only a few larger load reversals.
It is important to note that the duration of strong shaking is not the duration of the entire earthquake
event, which is much longer, as can be seen in acceleration time history data in Figure 2.8. While
there are a number of procedures that have been proposed to define the strong-shaking portion of
an acceleration time history or accelerogram, none have garnered the consensus of the seismologi-
cal or engineering communities. One such procedure for determining strong-motion duration is the
Parkfield 1966
El Centro 1940 - S00E
No. 2 - N65E
0.0
Helena 1935 - S90W Stone Canyon 1972
Melendy Ranch - N29W
Koyna 1967 - Long –0.5
PGA = 0.84g
PGA = 0.65g –1.0
0 10 20 30 40 50 60 70 80
Time (s)
FIGURE 2.8 Ground motions plotted using same horizontal and vertical scales. (Reproduced from Chopra,
A. K., Dynamics of Structures: Theory and Applications to Earthquake Engineering, 4th edition, Prentice-
Hall, Upper Saddle River, NJ, 2012. By permission of Prentice-Hall, Upper Saddle River, NJ.)
Engineering Seismology Overview 31
bracketed duration, which quantifies strong-motion duration as the time elapsed between the first
and last peak acceleration amplitudes that exceed a specified value (Bolt proposed 0.05g).
The dominant period at which the strong shaking occurs is often complementary and interrelated
to both PGA amplitude and strong-motion duration. Small PGA amplitude earthquakes cause little
damage to some structures because stresses may not be large enough, and even large PGA amplitude
earthquakes with short strong-motion duration may not cause enough stress reversals to cause any
significant structural damage. However, a relatively small PGA that continues for several seconds at a
uniform frequency can be devastating to some structures, particularly if the dominant period coincides
with the natural period of the structure, which results in a resonant condition, leading to severe damage.
This is the reason for the devastating damage experienced by some buildings during the 1985 Mexico
City earthquake, which had an epicenter approximately 400 km from Mexico City. The shaking in
poor soil areas was devastating to 10–20-story buildings because the frequency content was filtered as
the seismic waves entered these areas, leaving a dominant period of shaking of approximately 2 s (see
Figure 2.8). The dominant period of shaking coincided with the natural period of 10–20-story build-
ings, leading to a resonant condition; interestingly, other nearby buildings suffered little or no damage.
2.3.2 Site Characteristics
As discussed earlier, the strength of shaking usually diminishes as waves travel away from the focus
because of attenuation. Surface waves usually experience the same effect; this results in smaller
shaking for sites at greater distances from the epicenter. However, shaking can also strengthen
depending on the geologic conditions along the path of the wave between the focus and the site, as
well as the soil conditions and topography at the site. Soil condition and topography, in particular,
can greatly affect the characteristics of the input motion in terms of the amplitude and natural
period of the shaking at the site. A soft subsurface deposit or steep topography (ridge or hill) causes
the ground to vibrate like a flexible system, while a stiff bedrock causes the ground to vibrate
as a rigid body. That is, hard rock vibrates with the same frequency and amplitude as the input
motion, while the vibration of a flexible underlying deposit would depend on its inherent stiffness
and damping characteristics (in the same way as the dynamic response of a structure is affected by
its dynamic characteristics, as discussed in Section 2.3.3). For this case, the stiffness of the under-
lying material has the greatest effect on the characteristics of the input motion at a site (assuming
effects of damping to be negligible). Low-frequency (long-period) input motions are amplified more
at sites underlain by soft soils than those underlain by stiff ones, whereas stiff soil deposits amplify
high-frequency (short-period) input motions more than soft ones.
The inherent stiffness of the ground beneath a site is a function of its shear-wave velocity, and
thickness of the sediment above the bedrock. Shear-wave velocity is faster for hard rock than for soft
soil; consequently, seismic waves travel faster through hard rock than through soft soil. So when seis-
mic waves pass from rock into soil, they slow down. This slower speed must be accompanied by an
increase in wave amplitude (amplifying the ground shaking) in order for the flow of energy to remain
constant, which is required for the conservation of elastic wave energy. An increase in soft sediment
thickness has a similar effect (increased amplitude) because of a decrease in material density.
The most spectacular example of a deep flexible soil deposit amplifying low-frequency motions
occurred during the 1985 Mexico City earthquake. As discussed earlier (Section 1.1.3), a relatively
small rock acceleration caused a high PGA at the ground surface underlain by a very soft, high-
water-content mixture of clay and sand. The most severe damage was limited to an area near the city
center (the so-called Lake Zone), with little or no damage to buildings built on solid volcanic rock
just a few kilometers away (the so-called Foothill Zone); see Figure 1.8. The PGA for the Lake Zone
was about five times greater than that in the Foothill Zone. The frequency content for the two zones
was also very different; the Lake Zone had predominant periods of about 2 s, which was about 10
times larger than that in the Foothill Zone. Also, as shown in Figure 2.8, the strong shaking in the
Lake Zone continued for a long time. The duration of strong shaking along with the 2 s site period
32 Introduction to Earthquake Engineering
contributed to a peculiar damage pattern. Minor damage occurred to buildings of less than five sto-
ries and those of more than 30 stories; most other buildings collapsed or sustained extensive dam-
age. A simple rule of thumb for estimating the fundamental period of an N-story building is N/10 s
(Equation 3.19), which is the reason that buildings of about 20 stories collapsed; their fundamental
period matched that of the site ground motion period, creating a resonance condition. In effect,
these buildings suffered a double resonance condition: the amplification of the bedrock motion by
the soil deposit and the amplification of the soil motion by the structure.
During the 1989 Loma Prieta earthquake, the San Francisco Bay Area (some 100 km north
of the epicenter) experienced more damage than the epicentral region. The reason for the exten-
sive damage along the margins of the bay is because of the amplification of the PGA at soft-soil
sites. The bay basin is filled with an alluvial soil deposit of silts and clays, along with some layers
of sandy and gravelly soils. The upper deposit, or young mud, is composed of silty clay, known
as San Francisco Bay Mud. This material is of loose-to-medium density, and very compressible.
As shown in Figure 2.9, the ground stratum in the vicinity of the bay is divided into three zones
based on underlying material seismic response, from soft mud to hard rock. From Figure 2.9, it is
clear that the attenuation of seismic waves varied considerably in different regions of the bay due
to the varying subsurface soil conditions. The rock material attenuated the seismic waves relatively
quickly compared to the soft young bay mud found at the margins of the bay. The areas underlain by
this bay mud experienced the most damage because the soft mud greatly amplified ground motions.
These two examples clearly show that local site characteristics, particularly different underlying
ground types, greatly influence the seismic wave frequency content and amplitude, as will be shown
in Chapter 5. The National Earthquake Hazards Reduction Program (NEHRP) defines six ground
types based on shear-wave velocity in order to establish amplification effects for design purposes:
Type A has the largest stiffness and generates the smallest amplifications, while Type E is the
softest and generates the largest amplifications.
2.3.3 Structural Characteristics
Once seismic waves have been modified for ground conditions at the site of a structure, the struc-
ture responds to ground excitation based on its inherent characteristics, as well as the age of the
structure and the quality of its construction. The characteristics of a structure that have the largest
influence on its response include its weight and stiffness, and they will be covered in later chapters.
Damping can play a significant role when structures remain within the elastic limit, but remains a
much smaller effect for structures that deform into inelastic behavior as will be covered in Chapter
5. Although all these characteristics affect the displacement amplitude and natural period of a struc-
ture, stiffness has the greatest effect on structural response. A stiff structure vibrates with the same
frequency and amplitude as the input motion, while a flexible structure may or may not vibrate in
sync with the input motion, depending on how similar its natural period is to the input excitation
period. For instance, long-period (low-frequency) input ground excitations amplify the deforma-
tions of flexible structures more than those of stiff ones, whereas short-period (high-frequency)
input ground excitations amplify deformations of stiff structures more than those of soft ones. This
structural response will be discussed in detail in later chapters.
Engineering Seismology Overview 33
.53 .33
.45
SAN GREGORIO FAULT
.64 .33
Corralitos - CHAN1 : 0 deg .28
.47 EPICENTER .55
.50
.37
.54H
.60V .39H
Scale .56V .2
FIGURE 2.9 Horizontal PGA for the 1989 Loma Prieta earthquake (CSMIP = California Strong Motion
Instrumentation Program). (Reproduced from Chopra, A. K., Dynamics of Structures: Theory and Applications
to Earthquake Engineering, 4th edition, Prentice-Hall, Upper Saddle River, NJ, 2012. By permission of
Prentice-Hall, Upper Saddle River, NJ.)
The age and construction of structures play a major role in their performance during an
e arthquake. Older and/or poorly constructed structures are at a greater risk of failure during an
earthquake because of inadequate ductility, which is the ability of a structure to absorb the seismic
energy without failure. A structure can include seismic detailing to satisfy the ductility demand
during an earthquake, and to provide an adequate path for the seismic load to the foundation.
Addressing ductility demand and load path through proper detailing can allow a structure to per-
form adequately during even major earthquakes. The ductility demand can be addressed by allow-
ing the structure to behave in the inelastic range, provided that the stability of the overall system is
not compromised. An alternative approach to resisting seismic loads is to permit the movement of
structures; these systems are known as compliant systems.
34 Introduction to Earthquake Engineering
Seismic loads on structures result from inertial forces created by the ground accelerations. Based
on Newton’s second law (force equals mass times acceleration), the magnitude of these forces is a
function of the structure’s mass (or weight). Also, the acceleration is a function of the characteristics
of the input ground motion, particularly intensity, duration, and frequency content. The hazard a
given earthquake risk possesses to structures of different functions is accounted for by specifying
higher loads for structures of increased importance. Importance is classified in a range from unoc-
cupied buildings to critical structures such as hospitals and schools; as discussed in Section 8.4.2.
This method is too conservative because it does not properly account for uncertainties. For this
reason, the probabilistic method was developed.
The probabilistic approach is similar, but includes the uncertainties in each step. The probabilis-
tic approach, although quite complicated, has been incorporated in the various seismic hazard maps
used in contemporary seismic design codes. These codes have incorporated probabilistic seismic
hazard zone maps based on past geographic seismology to minimize risk to human life. These maps
depict where damage might occur, based on the probability of the occurrence of strong shaking and
the local geology (areas underlain by deep, loose sediments are at a greater risk for damage than
those directly over solid rock); see Figure 2.10. These maps can be used to draw probabilistic design
spectra, which will be discussed in detail in Chapters 5 and 8. With this information, structural
engineers can plan and design structures at a particular site using the methods covered in Chapter 8.
The key issue in the probabilistic approach found in the seismic design codes is the probability
of exceedance of a particular event for a specified intensity, over a certain period of time. The codes
consider three different earthquake levels for design: the maximum design earthquake (MDE),
which has a 10% probability of being exceeded over a period of 50 years; one with 5% probability
of being exceeded over a period of 50 years; and the maximum considered earthquake (MCE),
which has a 2% probability of being exceeded over a period of 50 years. Also, an earthquake
Engineering Seismology Overview 35
60
30
40° 40°
20
10
5
30° 30°
2
–120° –70°
–110° –80°
–100° –90°
Horizontal ground acceleration (%g)
with 2% probability of exceedance in 50 years
FIGURE 2.10 Probabilistic seismic hazard map showing ground shaking hazard zones. (From USGS
National Seismic Hazard Mapping Project website: geohazards.cr.usgs.gov.)
that has a significant chance of occurring during the life of a structure is known as the maximum
probable earthquake (MPE), and has a 50% probability of being exceeded over a period of 50 years.
Designing for the MCE provides the highest level of protection and is reserved only for the most
important or critical structures. These earthquakes can also be characterized in terms of the return
period (RP) using the following relationship:
−T
RP = c (2.8)
ln(1 − P )
where:
P is the probability of exceedance in T years
That is, the RPs in years for the three earthquake levels included in codes and the MPE are
as follows:
−T −50 years
MCE → RP = = = 2475 years
ln(1 − P ) ln(1 − 0.02)
−T −50 years
RP = = = 975 years
ln(1 − P ) ln(1 − 0.05)
−T −50 years
MDE → RP = = = 475 years
ln(1 − P ) ln(1 − 0.10)
−T −50 years
MPE → RP = = = 72 years
ln(1 − P ) ln(1 − 0.50)
36 Introduction to Earthquake Engineering
PROBLEMS
2.1 A Richter magnitude earthquake of 6.8 would radiate how much more energy than an
M L 5.8 earthquake?
2.2 A Richter magnitude earthquake of 5.0 is how much stronger than an M L 3.0
earthquake?
2.3 Estimate the local magnitude of a southern California earthquake recorded using a
standard Wood–Anderson seismograph that shows a trace amplitude of 23 mm, and the
P- and S-wave arrival times at the recording station of 07:19:45 and 07:20:09, respectively.
2.4 An earthquake at a transformed fault caused an average strike-slip displacement of
2.5 m over an area equal to 80 km long by 23 km deep. Assuming the rock along the
fault has the average shear stiffness of 175 kPa, estimate the seismic moment and
moment magnitude of the earthquake.
2.5 Estimate the seismic moment and moment magnitude of the February 27, 2010 Chile
earthquake. It is estimated that the thrust fault (the slip plane ends before reaching the
earth’s surface) caused an average strike-slip displacement of 5 m over an area equal
to 600 km long by 150 km deep. Assume the rock along the fault has the average shear
rigidity of 3 × 1011 dyne/cm2.
2.6 What is the energy difference between the Haiti and Chile earthquakes from
Example 2.2 and Problem 2.5, respectively?
REFERENCES
Bolt, B. A., Earthquakes, 5th edition, W.H. Freeman and Co., New York, 2004.
Chopra, A. K., Dynamics of Structures: Theory and Applications to Earthquake Engineering, 4th edition,
Prentice-Hall, Upper Saddle River, NJ, 2012.
Eberhard, M. O., S. Baldridge, J. Marshall, W. Mooney, and G. J. Rix, The MW 7.0 Haiti earthquake of
January 12, 2010; USGS/EERI Advance Reconnaissance Team Report: U.S. Geological Survey Open-
File Report 2010-1048, 58 p, 2010.
Kramer, S. L., Geotechnical Earthquake Engineering, Prentice-Hall, Upper Saddle River, NJ, 1996.
Richter, C. F., Elementary Seismology, W.H. Freeman and Co., San Francisco, CA, 1958.
Villaverde, R., Fundamental Concepts of Earthquake Engineering, CRC Press, Boca Raton, FL,
2009.
3 Single-Degree-of-Freedom
Structural Dynamic Analysis
After reading this chapter, you will be able to:
Traditional seismic design follows two approaches: one based on static analysis and the other
one based on dynamic analysis (see Chapter 8). The static analysis is formulated by estimating an
equivalent lateral seismic load as a fraction of the total system weight; the fraction is determined from
a notional earthquake, which is based on the risk posed by the seismicity of the site and is captured in
a single graph known as the response spectrum (to be discussed in Chapters 4 and 5). The dynamic
analysis procedure models the load effect of earthquakes more accurately by using either a set of spe-
cific earthquake ground motion time-history records (response time-history analysis procedure) or the
effect of the same notional earthquake as the equivalent static analysis (the design response spectrum),
but using a modal analysis (modal response spectrum analysis procedure). In the response time-history
analysis procedure, a linear or nonlinear mathematical model of the system is used to determine the
system response (displacements and accelerations, which are used to determine the internal loading in
each member) at each increment of time for a suite of ground motion acceleration time histories. With
the full history of the response, the absolute maximum internal loads can be determined by combin-
ing the results of the set of ground motion acceleration time histories; these loads can then be used to
design each member. The modal response spectrum analysis procedure uses established methods of
structural dynamics to determine system vibration mode shapes and their associated natural periods,
which along with the response spectrum are used to establish the maximum structural response. With
this maximum response, the maximum internal loads can be determined and used to design each
member. In this chapter, we explain the essential concepts of structural dynamics, also known as
vibration theory, needed to understand both dynamic seismic analyses procedures.
37
38 Introduction to Earthquake Engineering
u
u
m
sync, resulting in six DOFs. However, for our discussion in this chapter, the beam will be assumed
to be much stiffer than the columns, preventing rotations of the connections. Also, the two columns
move in parallel, and if we assume that the axial deformations of the columns and beam are rela-
tively small, we can reduce the entire system to a single-degree-of-freedom (SDOF) case as shown
in Figure 3.1. Finally, note that there is no damping in the system, the lack of which would cause the
system to move in perpetuity once set in motion. We will cover damping in Section 3.2.
+→ ∑ F = 0;
x − mu − ku = 0 ⇒ mu + ku = 0
where:
m is the mass of the system, discussed in more detail in Section 3.1.2
k is the lateral stiffness, discussed in more detail in Section 3.1.3
The double dot over the u indicates differentiation with respect to time, that is,
d 2u
u =
dt 2
The equation of motion, a second-order, linear, and homogeneous differential equation with
constant coefficients, can be written in the following form:
m mü
V=ku
where for convenience a new parameter is introduced, the natural circular frequency (from now on
called natural frequency) with units of radians per second (rad/s),
k
ωn = (3.2)
m
The solution to the second-order, linear, and homogeneous differential equation of motion is of
the form
u(t ) = eλt
This equation can be differentiated with respect to time to obtain the velocity,
u(t ) = λ eλt
and the velocity can be differentiated with respect to time to obtain the acceleration,
u(t ) = λ 2eλt
substituting the equations for displacement and acceleration into the equation of motion yields
λ 2eλt + ωn2eλt = 0
(λ2 + ωn2 )eλt = 0
In general, the exponential function is not zero, thus the quantity within parenthesis must be zero
in order to determine a valid solution to the differential equation,
(λ2 + ωn2 ) = 0
We then solve for the unknown parameter λ,
λ = ±ωni
where i is the imaginary unit of a complex number given as i = −1. As the equation of motion is a
second-order differential equation, two constants of integration are needed, A1 and A2,
This equation can be expressed in a polar form (in terms of sines and cosines) by making use of
Euler’s identities:
The result is
As the two trigonometric functions are real-valued solutions to the equation of motion, we can
express the general solution in real form as
which yields A = u(0). Next, we take a time derivative of Equation 3.3 and substitute the initial
0), in order to determine the constant B,
velocity (at time t = 0), u(
u(0)
u(t ) = u(0) cos ωnt + sin ωnt (3.4)
ωn
A graph of the solution is shown in Figure 3.3. This motion is described as harmonic (and there-
fore periodic) because it is a function of sine and cosine of the same frequency, ωn. The time required
to complete a full cycle (2π) is known as the natural period of vibration, Tn (s), and is determined as
(see Figure 3.3)
2π
ωnTn = 2π ⇒ Tn = (3.5)
ωn
. Tn = 2π /ωn
u u(0)
uo uo
b
a b c d e
uo
u(0)
a c e
t
β/ωn
The natural cyclic frequency, f n (hertz or cycles/s), is defined as the reciprocal of the natural
period of vibration, Tn (s), and is proportional to the natural circular frequency, ωn (rad/s), that is,
1 ω
fn = = n (3.6)
Tn 2π
Note that these three quantities are related and essentially represent the same physical quantity
expressed in different units.
An alternate formulation for the free vibration response of an undamped SDOF leads to the defi-
nition of two additional parameters introduced in Figure 3.3: the phase angle, β, and the amplitude,
uo. We start with Equation 3.3 and assume that constants A and B are related to new variables C and
β as follows:
A2 + B2 = C 2 (cos2 β + sin 2 β )
B C sin β
= = tan β
A C cos β
or
B u(0)
β = tan−1 = tan−1
A u(0)ωn
Substituting Equation 3.7 into Equation 3.3 yields the equation of motion in terms of the ampli-
tude and phase angle, that is,
u(t ) = A cos ωnt + B sin ωnt = C cos β cos ωnt + C sin β sin ωnt
1 1
cos β cos ωt = cos(ωt − β ) + cos(ωt + β )
2 2
1 1
sin β sin ωt = cos(ωt − β ) − cos(ωt + β )
2 2
42 Introduction to Earthquake Engineering
.
C = u(0)2 + (u(0)/ωn)2
β u(0)
α
.
u(0)/ωn
u(t ) = C cos(ωnt − β )
Similarly,
u(t ) = C sin(ωnt + α)
u(0)ωn
α = tan−1
u(0)
The two phase angles and the amplitude are related as shown in Figure 3.4.
3.1.2 Structural Weight
As noted earlier, the circular frequency is proportional to the mass, which in turn is directly pro-
portional to the weight. For seismic analysis of structures, the weight is defined as the total effective
weight W, which is specified in ASCE/SEI (2010) “Minimum Design Loads for Buildings and Other
Structures” standard, (from here on referred to as ASCE-7) as
where:
DL is the dead load of the structural system that is tributary to each floor
StL is the storage load, which is a live load in areas used for storage, such as a warehouse
PL is the partition load, when applicable
WPE is the operating weight of permanent equipment
SL is the flat roof snow load when it exceeds 30 psf, regardless of roof slope
LaL is the landscape loads associated with roof and balcony gardens
For partition loading, ASCE-7 requires the given partition load or 10 psf when determining the
effective weight. This is only applicable where the location of partitions is subject to change (such as
buildings designated as “office building” occupancy), where ASCE-7 requires a partition live load
of 15 psf for designing individual floor members for vertical loads, if the floor live load is <80 psf.
For information about estimating the other loads, see ASCE-7 specifications.
Once the dead load is estimated, one can determine the tributary weight of a floor or roof as
shown in Figure 3.5 for a multistory building. The weight of each floor and the tributary wall
loads halfway between adjacent floors is assumed to be concentrated or lumped at each floor level
(see Figure 3.6). The story weight, Wx, for this diagram is given as
Parapet
Roof W2
Wall and floor DL
tributary to the 2nd floor
of the building
h2
h2
2
1st level
h1 W1
2
h1 Tributary DL at this
level is assumed to be
concentrated at the
Base floor level and is
known as lumped
Wall and foundation DL
act directly on ground
Mid-height of C
Normal B
story above wall
A
WC
Shear
WA E wall
WD
WE D WB
Floor
Shear
wall
ke f
ua o
hq on
rt ti
Mid-height of
ea irec
story below
D
Here WD includes weight of the structure floor, suspended ceiling, mechanical equipment (unless
taken separately as WE ), and when applicable 10 psf for partitions or the actual partition weight.
In general, foundation weight and half of the first-story wall weight are assumed to act directly on
the ground and are commonly omitted in seismic calculations (though common practice, this is not
an explicit provision in seismic design specifications).
EXAMPLE 3.1
The load pressures for dead, live, and snow loads for the rectangular (30 ft × 40 ft) warehouse
structure shown are provided below. Determine the effective weight at the roof.
Roof DL = 45 psf
Floor DL = 65 psf
Wall DL = 25 psf
Floor LL = 250 psf (storage)
SL = 50 psf
44 Introduction to Earthquake Engineering
Solution
1. Identify the effective height of the roof in order to determine the tributary height of the
walls needed to calculate the effective weight of the walls.
In this case, the tributary height of the walls is the sum of the height from the roof to
midway to the base floor and the entire height of a parapet wall above the roof level as
shown in Figure E3.1:
h1 16 ft
htributary = + hparapet = + 4 ft =12 ft
2 2
( 40 ft ×12 ft )
WN– S,Walls = AreaN– S,Walls × Wall DL = ( 2 walls) ×25 psf = 24,000 lb
wall
(30 ft ×12 ft )
WE– W,Walls = ( 2 walls) × 25 psf =18,000 lb
wall
The weight of the roof level is calculated by multiplying the roof area with the roof
dead load.
This problem specifies a snow load pressure which is necessary for the seismic weight
calculation of the roof level. Based on Equation 3.7, snow loads exceeding 30 psf are
factored by 0.2 in the seismic weight calculation.
3. The effective seismic weight for the roof can be determined by summing the weight
contribution from each of the components.
As the roof is the only level above the base, the effective seismic weight of the roof is
denoted as W1:
4 ft
Roof htributary W
16 ft
Base
FIGURE E3.1 One story schematic for effective weight calculations for an idealized structural model.
Single-Degree-of-Freedom Structural Dynamic Analysis 45
EXAMPLE 3.2
Given the following rectangular (30 ft × 40 ft) two-story warehouse and loading shown, deter-
mine the effective weight at the roof and first floor:
Roof DL = 45 psf
Floor DL = 65 psf
Wall DL = 25 psf
Floor LL = 250 psf (storage)
SL = 50 psf
Solution
1. Identify the effective height of the first floor and the roof in order to calculate the effec-
tive weight of the walls at each level.
The effective height of the first level is the sum of the heights halfway to the adjacent
levels as shown in Figure E3.2:
h1 h2 16 ft 14 ft
heffective,1 = + = + =15 ft
2 2 2 2
The effective height of the roof is the sum of the heights from the roof midway to
the first level and the entire height of the parapet wall above the roof floor as shown in
Figure E3.2:
h2 14 ft
heffective,roof = + hparapet = + 2 ft = 9 ft
2 2
2. Calculate individual weights of components associated with each level of interest.
First level component weights:
For the first level, the seismic weight includes dead loads from the walls and floor,
whereas the live load is a result of storage load.
Weight of walls parallel to the N–S direction,
( 40 ft ×15 ft )
WN– S,Walls,1 = AreaN– S,Walls,1 × Wall DL = ( 2walls) × 25 psf = 30,000 lb
wall
2 ft
Roof W2
heffective, roof
14 ft
heffective, 1
1st level
W1
16 ft
Base
FIGURE E3.2 Two story schematic for effective weight calculations for an idealized structural model.
46 Introduction to Earthquake Engineering
(30 ft ×15 ft )
WE– W,Walls,1 = ( 2 walls) × 25 psf = 22,500 lb
wall
The floor load components on the first level include the floor dead load and the stor-
age live load,
This problem specifies a storage live load pressure which is necessary for the seis-
mic weight calculation of the first level. Based on Equation 3.7, storage live load, StL, is
factored by 0.25 in the seismic weight calculation.
WFloor LL,1 = AreaFloor,1× 0.25× StL = (30 ft × 40 ft ) × 0.25× 250 psf =75,000 lb
( 40 ft × 9 ft )
WN– S,Walls,roof = ( 2 walls) × 25 psf =18,000 lb
wall
(30 ft × 9 ft )
WE– W,Walls,roof = ( 2 walls) × 25 psf =13,500 lb
wall
The roof floor load components include the floor dead load and the snow load,
3.1.3 Structural Stiffness
To completely characterize the equation of motion for a linear elastic SDOF system, we need the
lateral stiffness, k. This variable, along with the mass, is needed to determine the circular frequency,
Single-Degree-of-Freedom Structural Dynamic Analysis 47
which once obtained can be used to determine the natural period of vibration of the system. The
stiffness is defined as the force or moment that results in a unit displacement or rotation at a DOF.
EXAMPLE 3.3
Determine the stiffness for a simply supported beam with a concentrated load F applied at mid-
span. This loading can be considered lateral with respect to the axis of bending.
Solution
1. Determine the deflection at point of interest (or DOF) for the given loading.
The deflection of the beam can be determined with one of the many structural analy-
ses methods (e.g., double integration, conjugate beam, virtual work, and moment-area)
used to c ompute deflections. Also, beam deflections can be obtained from deflection
tables such as those found in Part 4 of the American Institute of Steel Construction
Manual. The midspan deflection for a simply supported beam with a concentrated load
F is shown in Figure E3.3.
2. Determine the stiffness for the given loading and associated DOF.
The stiffness k is the ratio of force (or moment) to displacement (or rotation). The mid-
span deflection can be rearranged to determine the stiffness,
48EI
k=
L3
where the quantity EI is the flexural stiffness given by the product of the modulus of elas-
ticity, E, and the second moment of the cross-section about the axis of bending (moment
of inertia), I.
Following the approach of Example 3.3, we can derive equations for k for a number of simple
structural systems that can be modeled as SDOF (see Table 3.1 for a list of cases).
EXAMPLE 3.4
A fixed–fixed beam of length, L, with a concentrated weight, W, applied at midspan has a constant
flexural rigidity, EI. Idealize the beam and weight shown in Figure E3.4 into an SDOF system and
determine its natural and cyclic frequencies.
Solution
1. Idealize the structural system.
In order to idealize or model this structural system as a SDOF system, we assume that
the weight of the beam is small relative to the concentrated weight, W. In addition, we
limit the motion of the weight to vertical displacements, u, resulting in the lumped mass
model shown in Figure E3.5.
2. Determine the stiffness parameters.
Considering the displaced shape of the SDOF system as shown in Figure E3.6, the
static load–displacement relationship can be used to determine the stiffness associated
with the DOF.
F
FL3
Δ ∆=
48EI
FIGURE E3.3 Maximum deflection of a simply supported beam with concentrated force at
midspan.
48 Introduction to Earthquake Engineering
TABLE 3.1
Equivalent Stiffness Constants, k
Case Max Deflection, Δ Stiffness, k
Case 5wL4 384 EI
w 384 EI 5L3
Δ
F FL3 192 EI
Δ 192 EI L3
w wL4 384 EI
384 EI L3
Δ
Axially loaded bar F Fh AE
Δ
AE h
h
Springs in parallel F k1 + k2
k1 + k2
F
Springs in series F F 1 1 kk
+ 1/ + = 1 2
k1 k2 k k k + k2
F 1 2 1
m
u
FIGURE E3.5 Idealized lumped mass model of a fixed–fixed beam with concentrated weight at midspan.
FIGURE E3.6 Displaced shape of a fixed–fixed beam with concentrated force at midspan.
Single-Degree-of-Freedom Structural Dynamic Analysis 49
FL3
∆=
192EI
The stiffness of the SDOF system is determined by applying the same process as in
Example 3.3:
F 192EI
k= = 3
∆ L
u + ωn2u = 0
where the natural frequency is determined using Equation 3.2. Substituting for stiffness
and mass in terms of known quantities yields
The natural frequency, ωn, results in units of radians per second. In order to determine
the cyclic frequency, f, the units are converted from radians per second into cycles per
second by applying the conversion of 1 cycle = 2π radians or using Equation 3.6:
1cycle 4 3EIg
f = ωn = cycles/s or hertz
2π rad π WL3
EXAMPLE 3.5
Given a water tank supported on a slender column as shown in Figure E3.7, determine the natural
frequency and period of the system.
Solution
1. Idealize the structural system.
Similar to Example 3.4, the weight of the column is considered negligible with respect
to the weight of the water tank. Considering only the h
orizontal displacement of the
water tank, the lumped mass model is shown in Figure E3.7.
W = 10 kips Go
Tigers! W
E = 3 × 107 psi 50 ft
I = 20,000 in4
FIGURE E3.7 Water tank supported on a slender column and an idealized structural model.
50 Introduction to Earthquake Engineering
k 8333 lb/in
ωn = = = 17.94 rad/s
m (10, 000 lb/ (32.2 ft/s2 ×12 in/ft))
2π 2π
Tn = = = 0.35 s
ωn 17.94 rad/s
EXAMPLE 3.6
The space-grid roof structure may be considered rigid, and it has a dead load of 20 psf. The side
sheathing has a dead load of 10 psf. All steel columns are W10 × 30 (I = 170 in4). Determine the
dynamic properties, natural frequency, and period, in the E–W direction. Note that the first pair of
columns in Figure E3.8 is fixed–pinned, the second is pinned–pinned, and the last pair is fixed–
fixed. A plan view of the structure is shown in Figure E3.9.
Solution
1. Idealize the structural system into an SDOF system.
The frame can be modeled as an SDOF system assuming an rigid roof and the total
stiffness of the SDOF system is the sum of the individual column stiffnesses.
A B C
4′
12′
A B C
1
N
40′
2 @ 40′
3EI
kA1 = kA 2 =
h3
kB1 = kB2 = 0
12EI
kC1 = kC 2 =
h3
The total stiffness of the SDOF system is the sum of column stiffnesses in the E–W
direction.
2π 2π
Tn = = = 0.43 s
ωn 14.74 rad/s
Lateral force-resisting systems are used to resist lateral seismic (or wind) loads and are generally
categorized into one of three different categories: unbraced frames, braced frames, or shear walls.
Following are simplified derivations of the equivalent lateral stiffness constants for single-story
systems of each of these three cases.
Unbraced frames have rigid connections between columns and beams and carry lateral load by
developing moments in the columns and beams. If we consider the portal frame introduced earlier
52 Introduction to Earthquake Engineering
in this chapter (see Figure 3.1), we can derive the equivalent stiffness of the frame by assuming that
each column contributes equally to the frame stiffness. For example, for a portal frame with fixed–
fixed columns (such as the one listed in the sixth row of Table 3.1), which would be equivalent to a
frame with a very stiff beam, the total stiffness is given as
12 EI 24 EI (3.9)
K = (2 ) = 3
h3 h
where h is the height of the frame and EI is the flexural stiffness of the columns, which was defined
in the sixth row of Table 3.1. Silva and Badie (2008) gave the following formula for a general case
with variable beam-to-column stiffness ratio of α = Ib/Ic (assuming variable moments of inertia, but
constant modulus of elasticity) and variable beam-span-to-column-height ratio of κ = Lb/h:
24 EI 6α + κ
K= (3.10)
h3 6α + 4κ
For a portal frame case with pinned supports and a very stiff beam, K is given by the contribution
of two pinned–fixed columns (such as the one listed in the fifth row of Table 3.1) as
3EI 6 EI
K = (2 ) = 3 (3.11)
h3 h
Silva and Badie (2008) also gave a general case with variable beam-to-column stiffness ratio of
α and variable beam-span-to-column-height ratio of κ:
24 EI α
K= (3.12)
h3 4α + 2κ
EXAMPLE 3.7
The bridge systems depicted below has a rigid deck that weighs 500 kips. The lateral force-resist-
ing system consists of three concrete (E = 3000 ksi) columns as shown in Figure E3.10. Determine
the natural period of the system.
Solution
1. Idealize the structural system into a SDOF system.
The frame can be modeled as an SDOF system, and the total stiffness of the SDOF
system is the sum of the individual column stiffnesses.
Fixed–fixed Fixed–fixed
I = 15,000 in4 I = 15,000 in4
20 ft
Fixed–pinned
I = 15,000 in4
Soft soil
Rock
m 1.294
Tn = 2π = 2π = 0.762 s
K 87.9
Braced frames can be assumed to behave as cantilever beams that carry the lateral load by devel-
oping an internal shear force (known as panel shear) and a bending moment (known as overturning
moment). For portal braced frames similar to the one shown in Figure 3.7, we can assume that the
panel shear is carried by the diagonal members (one axial compression and the other axial tension),
whereas the overturning moment is carried by the axial forces in the columns. This braced frame
can be modeled as a statically indeterminate truss in which case the number of unknown reactions,
r, and unknown member forces, b, exceeds the number of equations of equilibrium (two times the
number of joints), 2j. To solve for the additional unknown in Figure 3.7, we can use compatibil-
ity of displacements. However, because the member sizes are initially unknown, we cannot apply
h F j = 4 ⇒ 2j = 8
2j < b + r
b=5 b+r=5
r=4
h
sin θ =
h2 + d2
θ d
cos θ =
h + d2
2
displacement compatibility. Therefore, we must make some simplifying assumptions to render the
truss statically determinate and perform an approximate analysis. First, the diagonal members can
be assumed to be stiff (in which case the diagonals can carry both tension and compression) or they
can be assumed to be slender (in which case the member in axial compression cannot support axial
force because it buckles). In the case of stiff diagonal members, we assume that the tension and
compression diagonals each carry half of the panel shear, whereas in the case of slender diagonals,
we assume that the panel shear is resisted entirely by the tension diagonal. The assumption of slen-
der diagonal members is used to derive the equation for the stiffness of a braced portal frame using
the principle of virtual work in Example 3.8.
The principle of virtual forces can be used to establish the displacement of a truss joint in any
direction. The process entails determining the internal forces in each member caused by the applied
loads and determining the virtual internal forces caused by a unit virtual load applied at the joint in
the direction in question. With these internal loads, the contribution from each member to the truss
deflection is summed to determine the total displacement,
m
∑n E A
N i Li
∆= i (3.13)
i i
i =1
where:
Ni is the axial force in each member caused by the actual loading
ni is the axial force in each member caused by a unit virtual load applied at the joint in the
direction in question
Li is the length of each member
Ai is the area of each member
Ei is the modulus of elasticity of each member
EXAMPLE 3.8
Derive an equation for the stiffness for a braced frame (see Figure 3.7) when the diagonal
members are slender.
Solution
1. Apply equilibrium to determine internal forces for actual and virtual loading.
Slender diagonals cannot carry compression load; thus, from statics we can deduce
that the left-hand column and beam are zero-force members, leaving us with the cases
shown in Figure E3.11 for both the actual and virtual loading.
Use the method of joints to determine the internal forces in the remaining members
by first drawing an FBD of the upper right-hand joints and applying equilibrium to the
resulting concurrent force systems as shown in Figure E3.12.
2. Apply virtual work (Equation 3.13) to determine the displacement at the top of the frame,
Δ. As the axial displacement of the column contributes little to the lateral d
isplacement
of the frame, it’s customary to ignore it:
∆= ∑ n
EA
=
E1A1
=F
d 2E1A1
3. Determine the stiffness (ratio of force to displacement) of the frame. The frame lateral
deflection equation can be rearranged to determine the stiffness as
d 2E1A1 (3.14)
k=
(h + d 2 )3 / 2
2
Single-Degree-of-Freedom Structural Dynamic Analysis 55
Δ Δ
F 1
θ θ
FIGURE E3.11 Actual and virtual loading with slender diagonal elements assumed.
N1 F 1
n1
N2 n2
Equilibrium: Equilibrium:
ΣFx = 0; F – N1 cosθ = 0 ΣFx = 0; 1 – n1 cosθ = 0
N1 = F/cosθ n1 = 1/cosθ
ΣFy = 0; – N2 – N1 sinθ = 0 ΣFy = 0; – n2 – n1 sinθ = 0
N2 = – F tanθ n2 = – tanθ
FIGURE E3.12 FBDs of joints due to the actual loading (left) and virtual loading (right).
where the quantity EA is the axial stiffness given by the product of the modulus of elas-
ticity, E, and the cross-sectional area, A.
EXAMPLE 3.9
The platform shown in Figure E3.13 is used at a stadium to film football games and is experiencing
large dynamic motions. Preliminary investigations indicate that the natural period is 0.9 s. Camera
personnel recommend the period to be limited to 0.3 s. Using steel (E = 29,000 ksi), determine the
required diameter of a system of diagonal ties (wires) to retrofit the system to the new specifications.
Solution
The system can be idealized as an SDOF system. Also, as the platform is square, only one
direction of motion needs to be considered, which is retrofitted with two diagonal ties (one along
each side—dash lines in Figure E3.13).
1. Determine the existing stiffness of the system using the observed natural period. In terms
of mass,
8 ft
8 ft
4.5 kips
6 ft
and the existing natural period of the SDOF system (Tn = 0.9 s), we can determine the
available stiffness using Equation 3.5:
2
m 2π
Tn = 2π ⇒ k = m
k Tn
That is,
2
2π 2π 2
k = m = 0.01164 = 0.5676 kips/in
Tn 0.9
2. Determine the required stiffness necessary to correspond to the limiting period. We now
determine the required stiffness to deduce the natural period to 0.3 s.
2
2π 2π 2
knew = m = 0.01164 = 5.108 kips/in
Tn 0.3
3. Determine the required diameter of diagonal ties needed to increase the stiffness of
the system. Determine the difference between the available and the new required
stiffnesses. This must be provided by the remedial diagonal ties. The change in s tiffness
can be applied to Equation 3.14 to determine the required radius of the diagonal ties:
d 2EA
knew − k = 2
(h2 + d 2 )3 / 2
Thus, the required diameter is 2r = 0.137 in, which can be provided by a 3/16-inch wire.
Single-Degree-of-Freedom Structural Dynamic Analysis 57
Shear walls can be assumed to behave as deep cantilever beams that carry the l ateral load by
developing internal shear force and bending moment as shown in Figure E3.14. That is, a shear
wall has deflections that are affected by both moment and shear (unlike regular beams, where
most of the deflection is caused by flexure). We can once again use the principle of virtual work
to establish the lateral displacement of the top of a shear wall. The process entails determin-
ing the internal shear and moment functions (diagrams) caused by the applied lateral load and
determining the virtual internal shear force and moment functions (diagrams) caused by a lateral
unit virtual load also applied at the top of the wall. With the internal shear force and moment
functions, the deflection of the top of the wall can be determined using the following virtual
work equation:
L L
M vV
∆=
∫ m f f dx +
EI ∫ κ GA dx (3.15)
0 0
where:
mf is the internal virtual moment caused by an external virtual unit load
Mf is the internal moment caused by the lateral load
v is the internal shear force caused by an external virtual unit load
V is the internal shear force caused by the lateral load
L is the length of the member
I is the moment of inertial of the member
E is the modulus of elasticity of the member
A is the area of the member
G is the shear modulus of elasticity of the member
κ is a constant that takes the following values depending on the cross-section of the shear wall:
κ = 1.2 for rectangular beams
κ = 10/9 for circular sections
κ = A/Aweb (ratio of total area to area of web) for I or box sections
EXAMPLE 3.10
Derive an equation for the stiffness for the cantilever shear wall shown in Figure E3.14.
Δ
F
FIGURE E3.15 Internal shear and bending moment diagrams for actual loading and virtual loading.
Solution
1. Apply equilibrium to determine shear force and bending moment functions for actual
and virtual loading.
Use structural analysis to determine the shear force and bending moment diagrams for
a cantilever beam, for both real load and virtual loading; see Figure E3.15.
2. Determine the displacement, Δ, at the top of the wall.
Apply virtual work (Equation 3.15) to determine the displacement at the top of the
wall, Δ. As the moments vary lineally and the shears are constant, we get the following
result:
L L
1 κ 1 κ
∆=
EI ∫ 0
mf ( x )Mf ( x ) dx +
GA ∫ v(x)V (x) dx = EI3 (h)(Fh)h + GA (1)(FF) h
0
3
Fh κFh
= +
3EI GA
3. Determine the stiffness (ratio of force to displacement) of the wall.
The wall lateral deflection equation can be rearranged to determine the stiffness,
which for a rectangular (κ = 1.2) shear wall is
3EI AG
K= + (3.16)
h3 1.2h
where the quantity EI is the flexural stiffness given by the product of the modulus of elasticity, E,
and the moment of inertia of the area, I, and AG is the shear stiffness given by the product of the
shear modulus of elasticity, G, and the area, A.
Figure 3.8 illustrates the relationship between the shear wall aspect ratio, d/h, and relative effect
of shear displacement, Δ/ΔM. ΔM is the displacement cause by moment only. Note that as the shear
wall gets taller, the effect of the shear force decreases rather quickly. For d/h < 1/5, the shear
Δ
Δ/ΔM
h
2
Δ d
= 1 + 0.94
ΔM h
1 d/h
0
FIGURE 3.8 Effect of shear force in the lateral deflection of a shear wall.
Single-Degree-of-Freedom Structural Dynamic Analysis 59
effect represents <4% so it may be neglected; h owever, for d/h > 1/3, the shear effect is important
as it is >10%.
We can also derive the equation for a shear wall that is fixed at the top and bottom (see Figure 3.9),
12 EI AG
K= + (3.17)
h3 1.2h
EXAMPLE 3.11
The structural system depicted in Example 3.6 (repeated here as Figure E3.16 for convenience)
has a roof weighing 20 psf and side sheeting weighing 10 psf. Assuming the lateral force-resisting
system in the N–S direction (one on each side) consists of ½-inch plywood (E = 1000 ksi and
G = 500 ksi) shear walls, determine the building’s natural period in this direction.
Solution
1. Determine the weight (proportional to the mass) of the SDOF system (same as
Example 3.6).
The weight of the SDOF system includes the roof dead load of 20 psf and the wall
sheathing weight around the perimeter of the building (240 ft). The weight due to
sheathing includes the tributary height of the roof, 10 ft, in the area calculation:
W = 20 psf (40′ × 80′) + 10 psf (10′ × 240′) = 88, 000 lb = 88 kips
A B C
1
N
40′
2 @ 40′
4′
12′
I = bh3 /12 = 0.5 in( 20 ft ×12 in/ft )3 /12 = 576, 000 in4
Ta = Ct hn x (3.18)
where:
hn is the height in feet above the base to the highest level of the structure
Ct and x are determined from ASCE-7 (Table 12.8.2; see Table 3.2)
An even more approximate relation that can be used to quickly check the values determined
using Equation 3.18 or the full dynamic analysis period is
Ta = 0.1N (3.19)
TABLE 3.2
Values of Approximate Period Parameters
Ct and x (ASCE-7 Table 12.8.2)
Structure Type Ct x
Steel moment-resisting frames 0.028 0.8
Concrete moment-resisting frames 0.016 0.9
Eccentrically braced steel frames 0.03 0.75
All other structural systems 0.02 0.75
Single-Degree-of-Freedom Structural Dynamic Analysis 61
0.0019
Ta = hn (3.20)
Cw
where:
100
x
hn 2
∑ Ai
CW =
AB hi [1 + 0.83(hi /Di )2 ]
i =1
∑ F = 0;x − mu − ku + p(t ) = 0 ⇒ mu + ku = p(t )
(3.21)
For the purpose of this analysis, we use a periodic, harmonic excitation force, which is a force
with magnitudes represented by sine or cosine as functions of time. (This force can also represented
a time-dependent displacement excitation, such as one produced by rotating machinery.) Sine and
cosine functions can also be applied to nonharmonic loadings with the response being obtained
using a Fourier method—a superposition of individual responses to the harmonic components of the
external excitations. First, assume a forcing function of the following form:
where:
po is the peak magnitude of the force
ω is the frequency of the force in radians per second
u Free-body diagram
p(t)
m p(t) m u
k V=ku
FIGURE 3.10 Free-body diagram of the SDOF model including a time-dependent force.
62 Introduction to Earthquake Engineering
This forcing function produces bound displacements provided that ω ≠ ωn = k /m , the resonant
condition, which will be discussed in more detail later in this section.
With a forcing function, we can solve the equation of motion following standard methods used in
the solution of nonhomogeneous differential equations. Equation 3.21 can be rewritten as
mu + ku = po sin ωt
The general solution to this equation can be expressed as a combination of the particular and
complementary solutions,
u(t ) = uc (t ) + uP (t ) (3.23)
where:
uc is the complementary solution, which is the solution to the homogeneous equation (the free
vibration solution) given by Equation 3.3 and repeated here for convenience,
uP is the particular solution, which usually takes the same form as the forcing function, that is,
u p (t ) = C sin ωt (3.24)
where C is the peak value, which is obtained by substituting Equation 3.24 into the original equation
of motion,
After factoring the sine function that appears in each term, and recognizing that this function is
not always zero, we can write this equation as
−mω 2C + kC = po
Solving for C,
po po (u )
C= = = st o
k − mω 2 k (1 − m /k ω 2 ) 1 − r 2
where:
r is the frequency ratio, that is,
ω
r=
ωn
We can rewrite the particular solution to include the phase angle, φ, as follows:
uP (t ) = uo sin(ωt − φ )
Single-Degree-of-Freedom Structural Dynamic Analysis 63
where:
uo is the maximum steady-state displacement response caused by the time-dependent
excitation force;
or
where:
Rd is the deformation response factor, which is the ratio of the maximum steady-state
displacement, uo, and the equivalent static displacement, (ust)o,
uo 1
Rd = = (3.26)
(ust )o 1 − r 2
0 ω < ωn
φ =
180 ω > ωn
Figure 3.11 shows graphs depicting the deformation response factor, Rd, and phase angle, φ, as
functions of frequency ratio, r. First, note that the displacement grows unbounded as r approaches
one, both from the right and the left. In actual practice, however, the system would come apart
or yield (changing the stiffness, thus changing the natural frequency). This point on the graph
corresponds to the resonant frequency, and as indicated earlier, the solution to the equation of
motion is not valid here. Also, the excitation force changes from being in phase with the natural
5
Deformation response factor, Rd
2
r= 2
0
180
uo is out of phase
Phase angle, φ
with force
degrees
90
uo is in phase
with force
0
0 1 2 3
Frequency ratio, r
FIGURE 3.11 Deformation response factor, Rd, and phase angle, φ, as a function of frequency ratio, r.
64 Introduction to Earthquake Engineering
vibration of the system to being out of phase; that is, the forcing function is additive up to r = 1;
and it begins to counter the force after r = 1. Furthermore, note that for small values of r (slowly
varying force), Rd is ∼1; and for large values of r (rapidly varying force), Rd approaches zero,
which corresponds to no displacement. An additional point of interest along the abscissa of these
graphs is r = 2, beyond which all values of Rd are ≤ 1. The point can be obtained by setting
Rd = 1.0. This implies that the dynamic displacement is always less than the static displacement
beyond r = 2 .
We now obtain the total dynamic response of the system by combining the complementary and
particular solutions,
po /k
u(t ) = A cos ωnt + B sin ωnt + sin ωt (3.27)
1− r 2
The two constants A and B are arbitrary and can be obtained by evaluating the equation at time
t = 0 (initial conditions). Again, the displacement at time t = 0 is u(0), that is,
po /k
u(0) = A cos(0) + B sin(0) + sin(0)
1− r 2
which again yields A = u(0). Next, differentiate the displacement equation with respect to time,
po /k
u (t ) = − Aωn sin ωnt + Bωn cos ωnt + ω cos ωt (3.28)
1− r 2
po /k
u (0) = − Aωn sin(0) + Bωn cos(0) + ω cos(0)
1− r 2
yields
u (0) po /k
B= − r
ωn 1− r 2
The complete solution is the forced vibration of an undamped SDOF system, which describes the
position of the mass as a function of time,
u (0) po /k p /k
u(t ) = u(0) cos ωnt + − r sin ωnt + o 2 sin ωt
2
(3.29)
ωn 1− r 1− r
A graph of which is given in Figure 3.12. In general, the resulting motion is the sum of periodic
motions of two different frequencies, ω and ωn, with different amplitudes. As noted earlier, when ω
approaches ωn, the motion becomes unbounded as t goes to infinity. However, when damping (dis-
cussed in Section 3.2) is included, the motion remains bounded.
p
Amplitude, po
Period, T = 2π/ω
2
Total response
Steady-state
response
1
u(t)/(ust)o
–1
–2
0 1 2
t/T
FIGURE 3.12 Forcing function and resulting response for r = 0.2, u(0) = 0.5po/k, and u (0) = ωn po /k .
energy (kinetic and potential) by transforming it into other forms of energy, such as heat. This type
of damping is inherent to a system. There may be cases where intentional damping is introduced
in a system, such as the shock absorbers in an automobile. Similar types of shock absorbers can
also be used for building systems, such as the fluid viscous dampers from Taylor Devices shown in
Figure 3.13. Similar to inertial and stiffness effects, damping effects can be characterized using a
force, which in this case is assumed to be proportional to the magnitude of the velocity and opposite
to the direction of motion. This type of damping is known as “viscous damping” and is the type of
damping produced in a body when restrained by surrounding viscous fluid. While viscous damp-
ing is not inherent in structural systems, this type of damping is much easier to manipulate in the
equation of motion than the more realistic damping from internal friction and is accurate enough
for most practical purposes.
FIGURE 3.13 Fluid viscous dampers (photographs courtesy of Taylor Devices, Inc.).
principle, where, in addition to the stiffness and inertial forces, a damping force proportional to the
velocity is included, as shown in Figure 3.14. In this case, we are using an oscillator model, which
is commonly used in structural dynamics textbooks (Chopra, 2012) because it is easier to visualize
the combined effects of damping, stiffness, and inertia.
Horizontal equilibrium of the FBD shown in Figure 3.14 yields the equation of motion,
Here, c is the damping coefficient, which is discussed in more detail in the next section.
Again, this is a second-order, linear, and homogeneous differential equation with constant
coefficients, the solution to which is also of exponential form, that is,
u(t ) = e ρ t
u(t ) = ρeρ t
and the velocity can be differentiated to determine the acceleration,
u(t ) = ρ 2eρ t
u u
k ku
m u
m
c cu
Viscous
damper
FBD
FIGURE 3.14 Idealized SDOF system and FBD for a portal frame.
Single-Degree-of-Freedom Structural Dynamic Analysis 67
substituting the equations for the displacement, velocity, and the acceleration into Equation 3.30
yields
(mρ 2 + cρ + k )eρ t = 0
In general, the exponential function is not zero, thus the quantity within parenthesis (also known
as the characteristic equation) must be zero,
mρ 2 + cρ + k = 0
−c + c 2 − 4mk
2 ρ1 =
−c ± c − 4mk 2m
ρ= ⇒
2m c 2 − 4mk
ρ = −c −
2 2m
Since the equation of motion (Equation 3.30) is a second-order differential equation, we need
two constants of integration, C1 and C2, which are determined from initial conditions (displacement
and velocity).
−c c2 − 4 mk − c2 − 4 mk
t
u(t ) = e 2 m C1e
t t
2m + C2 e 2m
This result leads to three different solutions depending on the value of the quantity under the
square root operator (see Figure 3.15):
1. If c2–4 mk > 0, exponents are real numbers. The motion for this case decays exponentially;
thus, no oscillation occurs. For this case, the system is classified as overdamped.
2. If c2–4 mk < 0, exponents are imaginary values, and oscillatory motion occurs. For this
case, the system is classified as underdamped.
3. If c2–4 mk = 0, then there is a boundary between oscillatory and nonoscillatory motion.
For this case, the system is classified as critically damped.
Structural systems are typically underdamped, so our discussion on damping will be focused
primarily on this type of system. However, critically damped systems are important because they
are used to define the damping ratio, ζ, which is typically used to describe damping in structural
systems. For a critically damped system, c2−4 mk = 0, which leads to the definition of the critical
damping coefficient, ccr:
68 Introduction to Earthquake Engineering
1.5
Critically damped, ζ = 1
1
Overdamped, ζ = 2
0.5
u(t)/u(0)
0
0 1 2 3
t/T
–0.5
Underdamped, ζ = 0.1
–1
–1.5
2k
ccr = 2 km = 2mωn = (3.31)
ωn
The other parameters were defined earlier. With the critical damping coefficient, ccr, we now
define the damping ratio or damping factor (which is an expedient way to define damping for practi-
cal structural systems and is usually expressed in a percentage form) as
c
ζ= (3.32)
ccr
We can also use this value to define the three different general states of damping: overdamped
systems have ζ > 1, critically damped systems have ζ = 1, and underdamped systems have ζ < 1, as
shown in Figure 3.15. Since most structural systems are underdamped, in this book we only provide
the solution for this case. Rewriting the unknown parameters ρ1 and ρ2 as
(
ρ 1, ρ 2= −ζ ± ζ 2 −1 ωn )
This will be complex as ζ < 1. We can write this in complex number form as
(
ρ 1, ρ 2= −ζ ± i 1 − ζ 2 ωn )
1−ζ 2 ωn t 1−ζ 2 ωn t
u(t ) = e−ζωn t C1ei + C2e−i
This equation can be expressed in polar form (in terms of sines and cosines) by making use of
Euler’s identities; and because trigonometric functions are real-value solutions to the equation of
motion, we can express the general solution in real form as
where A and B are arbitrary constants. While this motion is periodic, it is no longer of constant
frequency. In fact, the frequency (and the period) is a function of the damping ratio—the damped
frequency,
ω D = ωn 1 − ζ 2 (3.34)
2π 2π Tn
TD = = = (3.35)
ω D ωn 1 − ζ 2
1− ζ 2
We now use initial conditions (at time t = 0) to solve for constants A and B. First, the initial
displacement (at time t = 0) is u(0), that is,
which yields A = u(0). Next, differentiating the displacement equation with respect to time we
obtain
u(t ) = −ζωne−ζωn t [ B sin ωD t + A cos ωD t ] + e−ζωn t [ BωD cos ωDt + AωD sin ωD t]
where:
u(0) + ζωnu(0)
B=
ωD
Again, the complete solution is the free vibration response of a damped SDOF system, which
describes the position of the mass as a function of time,
u(0) + ζωnu(0)
u(t ) = e−ζωn t sin ω t + u(0) cos ω t (3.36)
ωD D D
A graph of which is given in Figure 3.16. The damped structure clearly shows motion decay
over time. The decay in the motion is mathematically produced by the exponential part of the
solution.
This oscillation decay can be used to experimentally determine the inherent damping in a system.
The process entails setting the structure into free vibration from an initial displacement and obtain-
ing a record of the motion. The rate of decay of amplitude of motion is characterized by a ratio
of the value of two successive peak displacement amplitudes, which is known as the logarithmic
decrement, δ. The values of two successive peak displacement amplitudes are
u(0)
1 Undamped
u1
ρe–ζωnt structure
Damped
u2 structure
u(0)
u(t)
–ρe–ζωnt Tn = 2π/ωn
TD = 2π/ωD
FIGURE 3.16 Comparison of free vibration response of undamped and damped SDOF systems.
where ι is a constant that depends on the initial conditions. The ratio of these displacements is
u1 ιe−ζωn t1
= −ζωn ( t1 +TD ) = eζωn ( t1 +TD )−ζωn t1 = eζωnTD
u2 ιe
Taking the natural log of both sides, we can define the logarithmic decrement, δ,
u
δ = ln 1 = ζωnTD (3.37)
u2
u1
δ = ln = ζωnTD (3.38)
u2
In both cases, δ can be defined in terms of ζ only. This is accomplished by replacing the damped
period as follows:
ζωn 2π 2πζ
δ = ζωTD = = (3.39)
2
ωn 1 − ζ 1− ζ 2
For small damping ratio values (ζ < 0.2), the value under the square root operator is close to
unity and the logarithmic decrement can be approximated as
δ ≈ 2πζ (3.40)
Single-Degree-of-Freedom Structural Dynamic Analysis 71
Furthermore, for lightly damped systems where successive oscillation peaks have similar ordi-
nates, two nonconsecutive amplitudes ui and ui+n (where n is any integer) can be used to determine δ,
ui
ln = nζωnTD = nδ (3.41)
ui +n
Also, note that the equation of motion can be rewritten in the following form:
3.2.2 Structural Damping
As noted earlier, in addition to attenuating vibration amplitudes, damping has the effect of lower-
ing the natural frequency, and thus lengthening the period from the natural state. Most structural
systems possess small inherent damping, <20% as listed in Table 3.3; and as shown in Figure 3.17,
damping effects on the period are negligible when the damping ratio is low. Thus, the damped and
undamped periods are practically equal, and in practice, TD is usually taken as Tn (except when
supplemental damping is added to a structural system).
The damping ratio characterizes how fast oscillations decay from one cycle of motion to the
next and has a significant effect on the response of elastic systems. However, the inelastic response
is dominated by ductility, and damping has little attenuation effect relative to ductility—this will
be described in Chapter 5. For this reason, seismic design guidelines (which assume inelastic
behavior) assume a uniform 5% damping for all building structural systems. Furthermore, current
k nowledge of damping mechanisms in structural systems, particularly buildings, is rather limited,
with l ittle ongoing research in the area. It is believed that damping in building systems is the result
of the f ollowing damping mechanisms: friction in structural elements (friction in bolted and nailed
connections, and cracking of concrete), intrinsic material damping, friction in nonstructural com-
ponents and their connections to the structure, and soil–structure interaction (radiation of seismic
waves back into the soil and intrinsic damping in the soil) (Villaverde, 2009). All of the aforemen-
tioned damping mechanisms are difficult to characterize, thus making it difficult to determine
damping analytically for actual structures. Damping ratios tabulated in Table 3.3 (Lindeburg and
TABLE 3.3
Typical Damping Ratios
Type of Construction ζ
Steel frame with welded connections and flexible walls 0.02
Steel frame with welded connections, normal floors, and exterior cladding 0.05
Steel frame with bolted connections, normal floors, and exterior cladding 0.10
Concrete frame with flexible interior walls 0.05
Concrete frame with flexible interior walls and exterior cladding 0.07
Concrete frame with concrete or masonry shear walls 0.10
Concrete or masonry shear walls 0.10
Wood frame and shear walls 0.15
1
ωD Tn
= = 1 – ζ2
ωn TD
Tn
TD
0 ζ
0 1
McMullin 2008) are only provided to illustrate that real structures do not possess inherent damp-
ing >15%. From the given data, it should also be clear that the damping ratio depends on the type
of building construction.
EXAMPLE 3.12
A free vibration test is conducted on an empty elevated water tank. A cable attached to the tank
applies a lateral force of 16.4 kips and pulls the tank horizontally by 2 in. The cable is suddenly cut
and the resulting free vibration is recorded. At the end of four complete cycles, the time is 2.0 s,
and the amplitude is 1 in. From the given information, compute the damping ratio, natural period,
effective stiffness, effective weight, damping coefficient, and number of cycles required for the
displacement to decrease to 0.2 in.
Solution
1. Determine the dampin
As ui = u1 = 2 in at t = 0 and ui+n = u5 = 1 in at n = 4, from Equation 3.41
u 2
ln 1 = ln = nζ 2π
u5 1
1 2
Therefore, ζ = ln = 0.0276 or 2.76%.
4 × 2π 1
The low damping ratio allows for a small damping assumption.
2. The damped period is determined by dividing the lapse time by the number of cycles,
2s
TD = = 0.5 s /cycle
4 cycles
In addition, due to the small damping assumption, the natural period can be approxi-
mated as the damped period.
Tn ≅ TD = 0.5 s
Single-Degree-of-Freedom Structural Dynamic Analysis 73
3. Determine the equivalent stiffness of the water tank assuming a linear force–displacement
relationship
F 16.4 kips
k= = = 8.2 kips/in
u 2 in
4. Determine the effective weight using the period and the stiffness
2π 2π
ωn = = = 12.57 rad/s
Tn 0 .5 s
k 8.2 kips/in
m= = = 0.0519 kip ⋅ s2 /in
ωn2 (12.57 rad/s)2
( )
c = ς ( 2 Km ) = 0.0276 2 (8.2kips/in)(0.0519 kip ⋅ s2 /in) = 0.036 kip ⋅ s/in
6. Determine the number of cycles for displacement to reduce to 0.2 inches using the
logarithmic decrement δ
u
ln i = nζ 2π ⇒
ui +n
1 u 1 2 in
n= ln i = ln = 13.3 cycles or 13 cycles
276)2π 0.2 in
ζ 2π ui +n (0.02
For the purpose of this analysis, we again use a periodic, harmonic excitation force. The force
excitation can also be represented as a time-dependent displacement excitation as will be shown in
Section 3.2.4. We begin again by assuming a forcing function of the form
p(t ) = po sin ω t
where:
po is the peak magnitude of the force
ω is the frequency of the force in radians per second
This forcing function produces bound displacements; even for the resonant condition, pro-
vided damping is nonzero. Again, with this forcing function, we can solve the equation of motion,
74 Introduction to Earthquake Engineering
mu + cu + ku = po sin ω t , following standard methods used in the solution of nonhomogeneous
differential equations. The general solution to this equation is of the same form as that for the
undamped case,
u(t ) = uc (t ) + uP (t ) (3.44)
where uc is the complementary solution, which is the solution to the homogeneous equation for
the underdamped case (the free vibration solution) given by Equation 3.33, and repeated here for
convenience,
up is the particular solution, which usually takes the same form as the forcing function, that is,
where C and D are constants that are obtained by substituting the particular solution into the
original equation of motion,
po (1 − r 2 ) po 2ζ r
C= and D =−
k (1 − r 2 )2 + (2ζ r )2 k (1 − r 2 )2 + (2ζ r )2
po /k
uP (t ) = ((1 − r 2 )sin ωt − 2ζ r cos ωt ) (3.46)
(1 − r 2 )2 + (2ζ r )2
We can rewrite the particular solution to include the phase angle, φ, as follows:
uP (t ) = uo sin(ω t − φ)
where:
(ust )o
uo = C 2 + D 2 =
(1 − r 2 )2 + (2ζ r )2
2ζ r
φ = tan−1 (−D /C ) = tan−1
1 − r 2
uo 1
Rd = = (3.47)
ust (1 − r ) + (2rζ )2
2 2
Single-Degree-of-Freedom Structural Dynamic Analysis 75
Note that this ratio is a function of r and ζ only. Maximizing Rd in terms of r yields several
interesting conditions. First, take the partial derivative of Rd with respect to r is
∂Rd r (1 − r 2 − 2ζ 2 )
= =0
∂r [(1 − r 2 )2 + (2rζ )2 ]3/2
Figure 3.18 shows graphs depicting the deformation response factor, Rd, as a function of
f requency ratio, r. This graph is the same as that shown in Figure 3.11, but includes damp-
ing. Notice that in this case, the displacement does not grow unbounded as r approaches 1.
In fact, as damping increases the displacement at the resonant frequency decreases, and for
damping ratios exceeding 70.7%, the dynamic displacement is always less than the static dis-
placement. Again, for small values of r (slowly varying force), Rd is approximately equal to 1;
and for large values of r (rapidly varying force), Rd approaches zero, which corresponds to no
displacement.
We now obtain the total dynamic response of the system by combining the complementary
solution (transient response) and particular solution (steady-state response),
ζ = 0.01
5
4 ζ = 0.1
3
Rd
ζ = 0.2
2
1
ζ = 0.7
ζ=1
0
0 1 2 3
r
The two constants A and B are arbitrary and can be obtained using the initial (at time t = 0) con-
ditions. Again, the displacement at time t = 0 is u(0), which yields the following when substituted
into Equation 3.48:
(ust )o sin(−φ)
A = u(0) −
(1 − r 2 )2 + (2ζ r )2
A graph of the complete solution that describes the position of the mass as a function of time is
given in Figure 3.19 (similar to Figure 3.12, but with damping). In general, the resulting motion is
the sum of periodic motions of two different frequencies (ω and ωn) and amplitudes. Note that the
transient response (the one due to the inherent motion in the system) attenuates as t goes to infinity.
Also, as noted earlier for undamped systems, the motion becomes unbounded when ω approaches
ωn, as t goes to infinity. However, for damped cases, motion remains bounded to a maximum of 0.5ζ,
as shown in Figure 3.20.
The effect of a harmonic force excitation on the foundation of a supporting system can be deter-
mined as the magnitude of the force transmitted to the support through the spring (stiffness) and
the damping elements,
( fT )o = (ust )o Rd k 2 + c 2ω 2
2 Total response
Steady-state
response
1
u(t)/(ust )o
–1
–2
0 1 2
t/T
FIGURE 3.19 Harmonic force response for r = 0.2, ζ = 0.05, u(0) = 0, and u (0) = ωn po /k .
Single-Degree-of-Freedom Structural Dynamic Analysis 77
15
Steady-state
amplitude
1
10
2ζ
Envelope
curves
5
u(t)/(ust )o
–5
–10 – 1
2ζ
–15
0 2 4 6 8 10
t/T
Rewriting this as a fraction of equivalent static force, po, we get the transmissibility, Tr,
( fT )o
= Rd 1 + (2ζ r )2
po
or
( fT )o 1 + (2rζ )2
Tr = = (3.49)
po (1 − r 2 )2 + (2rζ )2
EXAMPLE 3.13
A sensitive, 50-lb instrument that requires insulation from vibration is being installed in a building
where a reciprocating machine is in use. The machine causes the building floor to vibrate in a
harmonic motion at frequency of 1000 cycles per minute. The equipment is going to be installed
on four springs of equal stiffness and negligible damping. Determine the stiffness of each spring if
the amplitude of the transmitted vibration is to be limited to <15% of the floor vibration.
Solution
After idealizing the system as an SDOF system, use transmissibility with ζ = 0 to determine the
required stiffness to limit the vibration transmitted.
1+ ( 2rζ )2 1
Tr = =
2 2
(1− r ) + ( 2rζ ) 2 ±(1− r 2 )
1. Determine the maximum required frequency ratio, r. Since Tr < 1, r > 2, so, apply the
negative sign throughout the denominator to obtain
1
Tr =
(r 2 − 1)
78 Introduction to Earthquake Engineering
and establish an inequality such that the transmissibility ratio is <0.15, that is,
1
< 0.15
(r 2 − 1)
solving this inequality for r, we determine that frequency ratio, r, must be >2.77.
2. Determine the required natural cyclical frequency. Use the frequency ratio in the follow-
ing form:
f
r= > 2.77
fn
16.67 Hz
> 2.77 ⇒ fn < 6.02Hz
fn
3. Determine the mass
4. We now determine the total required stiffness to limit the vibration transmitted.
1 kt 1
fn = < 6.02
2π m s
ut
u
u ut
k fs
m fI
c fD
ug
ut (t) = ug (t) + u(t) FBD
ug
FIGURE 3.21 Idealized SDOF system and free-body diagram for a portal frame.
Single-Degree-of-Freedom Structural Dynamic Analysis 79
of motion for the system by applying equilibrium to the FBD of the mass using D’Alembert’s prin-
ciple, where the stiffness and damping forces are proportional to the relative displacement and the
inertial force to the total mass displacement. Again, we are using an oscillator model because it is
easier to visualize the combined effects of damping, stiffness, and inertia.
Horizontal equilibrium of the FBD shown in Figure 3.21 yields the equation of motion,
∑ F = 0;
x f I (t ) + f D (t ) + f S (t ) = 0
where:
fI (t ) = mut (t ) is the inertial force
f D (t ) = cu(t ) is the damping force and fS (t ) = ku(t ) is the stiffness force
This equation can be further rewritten in terms of the relative and ground motion by substituting
the definition of total acceleration as ut (t ) = u(t ) + ug (t ) into the equation of motion,
The right-hand side is the effective support excitation loading that opposes the sense of ground
acceleration and is similar to the time-varying force discussed in Section 3.1.5.
Alternatively, we can write this equation in terms of the total displacement, in which case the
right-hand side is the effective loading; the resulting response is the total displacement of the mass
from a fixed reference, rather than displacement relative to the moving base:
ug (t ) = ugo sin ωt
where:
ugo is the peak amplitude of the ground displacement
ω is the frequency of the support motion in this case
The two terms on the right-hand side can be combined into an equivalent form by performing
a trigonometric transformation such as the one performed in Section 3.1.1 to obtain u(t) = C sin
(ωnt + α),
where:
cω
tan α = = 2rζ
k
Equation 3.53 is the differential equation for the oscillator excited by the harmonic force Fo sin
(ωt + α); the solution to which is of the same form as that for Equation 3.43. That is, following stan-
dard methods used in the solution of nonhomogeneous differential equations such as the solution in
Equation 3.44, which includes the complementary (transient) and particular (steady state) solutions.
The steady-state solution describes the relative transmission of support motion to the mass,
Fo /k sin(ωt + α − φ)
u t (t ) =
(1 − r 2 )2 + (2rζ )2
or
u t (t ) 1 + (2rζ )2 sin(ωt + α − φ)
= (3.54)
ugo (1 − r 2 )2 + (2rζ )2
Maximizing this relationship leads to the expression for transmissibility, which was discussed
earlier and given by Equation 3.49. That is, transmissibility of motion from the foundation to the
structure is given by the same function as the transmissibility of force from the structure to the
foundation. This is an important expression in vibration isolation; in this context, it is the degree
of relative isolation defined as the ratio of the amplitude of motion of the system to the static
displacement:
uot 1 + (2rζ )2 ut
Tr = = = o (3.55)
ust (1 − r 2 )2 + (2rζ )2 ugo
where:
ust = Fo/k is the static displacement
Again, we note that this is a function of the frequency ratio, r, and the damping ratio, ζ, only.
This can also be expressed in terms of acceleration,
uot 1 + (2rζ )2
Tr = = (3.56)
ugo (1 − r 2 )2 + (2rζ )2
Figure 3.22 shows graphs depicting the transmissibility as a function of frequency ratio,
r, for various values of ζ. This graph is similar to that shown in Figure 3.18.
From this graph, the following observations are noted:
3
ζ=0
2.5
ζ = 1/5
Transmissibility, Tr
2
ζ = 1/4
1.5
ζ = 1/3
ζ = 1/2
1
ζ = 1/2
ζ = 1/3
0.5
ζ = 1/4
ζ = 1/5
0
0 0.5 1 2 1.5 2 2.5 3
Frequency ratio, r
3. As r approaches zero (slowly varying displacements), uot = ugo; that is, the mass moves
rigidly with the ground motion.
4. r = 1 again leads to resonance; as ζ→0, Tr→∞.
5. Increasing damping when r > 2 decreases the effectiveness of the vibration isolation
system.
6. As r approaches infinity (rapidly varying displacements), uot = 0; that is, the mass stays still
while the ground beneath it moves.
1. Since Tr > 1 for r < 2 and Tr < 1 for r > 2, vibration isolation can be achieved only in
the range of r > 2 .
2. Since damping increases transmissibility in the isolation range (r > 2 ), the most effective
vibration absorbers consist of spring elements having little or no damping. This implies a
tradeoff when selecting a soft spring to reduce the transmitted force but increases the static
displacement or vice versa with a stiff spring.
uo r2
= (3.57)
ugo (1 − r 2 )2 + (2rζ )2
EXAMPLE 3.14
A 50-lb package is suspended in a box (damping of 2%) by two springs each having a stiffness
k of 250 lb/in. The box is transported on the bed of a truck, which due to the suspension motions
experiences vertical harmonic excitations of amplitude u(t) = 1.5 in sin (2 rad/s ⋅ t). The owner
of the box has specified that the maximum relative displacement of the package be <0.05 in.
Does the system satisfy this requirement?
Solution
There are two ways to solve this problem; we can directly determine the relative displacement or
use transmissibility to determine the total displacement and then compute the relative displace-
ment. In both cases, we first need the frequency ratio.
82 Introduction to Earthquake Engineering
Stiffness,
Natural frequency,
kt 500
ωn = = = 62.16 rad/s
m 0.1294
ω = 2 rad/s
ω 2
r= = = 0.0322
ωn 62.16
2. Determine the relative displacement of the mass. Using the transmissibility ratio defi-
nition of Equation 3.57, the relative displacement of the system can be determined.
ugo is the amplitude of the system displacement given in the displacement equation as
1.5 in,
uo r2
=
ug o (1− r )2 + ( 2rζ )2
2
ug o r 2 (1.5)(0.0322)2
⇒ uo = = = 0.00155′′ < 0.05′′
(1− r 2 )2 + ( 2rζ )2 (1− (0.0322)2 )2 + (2(0.0322)(0.05))
2
uot 1+ ( 2rζ )2
Tr = =
ug o (1− r 2 )2 + ( 2rζ )2
2
ug o 1+ ( 2rζ )2 (1.5) 1+ (2(0.0322)(0.05))
⇒ uot = = = 1.50155′′
(1− r 2 )2 + ( 2rζ)2 (1− (0.0322)2 )2 + ( 2(0.0322)(0.05))2
Same as above!
Single-Degree-of-Freedom Structural Dynamic Analysis 83
1. The dynamic effect can be replaced with a slowly applied force that produces deforma-
tion umax = uo (or acceleration ümax = üo), which is determined by a dynamic analysis. As
shown in the prior two sections, this displacement can equally be produced by a dynami-
cally applied force or a support excitation. Displacements can be used to determine an
equivalent, external static force, Fs,max, as shown in Figure 3.23, which can in turn be used
to determine all internal forces.
2. With known dynamic displacements at column ends, the forces (moment and shear) can
be determined using individual element stiffness via fixed-end moment analyses, see a
structural analysis textbook. Figure 3.24 depicts two cases we have used in this chapter.
The first case shows the base shear and moment, which are also equal to the maximum
internal shear and moment. The second case has maximum shear and moment at the top
and bottom of the column.
In both cases, maximum stresses (shear and flexural) can be determined using standard strength
of materials procedures. For example, the maximum shear stress is
VQ
τ=
It
uo
Fs,max
Tributary mass
Fs,max = k uo = m üo
Lateral stiffness
uo
uo
3EI V 12EI
Vb = kuo = uo M V = kuo = uo
h3 h3
EI EI
Mb = hVb = 3 uo M=6 uo
h2 h2
M
Vb V
Mb
FIGURE 3.24 Internal shear force and moment for pinned–fixed and fixed–fixed columns.
84 Introduction to Earthquake Engineering
where:
Q is the first moment of the area above the plane in question
I is the moment of inertia of the cross section
t is the width of the section at the plane in question
This equation yields the following maximum stresses for rectangular and I-shaped sections:
Rectangular section,
Vmax
τ max = 1.5 (3.58)
A
I-shaped section,
Vmax
τ max ≈ (3.59)
Aweb
where:
A is the area of the entire rectangular cross section
Aweb is the area of the web; in practice, this is equal to the product of the depth of section and
thickness of the web
M max
σmax = (3.60)
S
where:
S is the section elastic modulus
EXAMPLE 3.15
Given the building frame shown in Figure E3.17 with damping of 5% and load p(t) = 200 lbs sin 5.3 t,
determine (a) maximum displacement u, (b) the maximum base shear, and (c) the maximum normal
stresses in the columns (Ix = 75 in4 and Sx = 18 in3 for W8 × 21). Let us assume that the beam is rigid.
Solution
1. Determine mass, stiffness, and natural frequency of the SDOF system.
The building frame can be modeled as an SDOF system assuming that the horizontal
beam is rigid and only lateral deformations of the columns occur. The stiffness of the
system is then the sum of column lateral stiffnesses. The mass and stiffness of the SDOF
system are calculated as follows:
u
p(t) 15 kips
15 ft
Mass,
W 15 kips(1000 lb/kip)
m= = = 38.86 lb ⋅ s2 /in
g 386 in/s 2
Stiffness,
4
k=
3EI 3EI
+ = 2 3( 29, 000, 000 psi)(75 in ) = 2237.6 lb/in
L3 L3 3
(15 ft ×12 in/ft )
The natural frequency of the SDOF system is determined using Equation 3.2:
k 2237.6 lb/in
ωn = = = 7.59 rad/s
m 38.86 lb ⋅ s2 /in
po 200 lb
(ust )o = = = 0.08938 in
k 2237.6 lb/in
Also, given forcing frequency of ω = 5.3 rad/s, the ratio of forcing frequency to natu-
ral frequency can be determined as follows:
ω 5 .3
r= = = 0.6984
ωn 7.59
Now substituting this value into Equation 3.47, the maximum steady-state
displacement is
(ust )o 0.08938 in
uo = = = 0.1729 in
(1− r 2 )2 + ( 2rζ )2 (1− 0.6982 )2 + ( 2(0.698)(0.05))2
3. Maximum base shear.
Calculate maximum base shear at the support for each column using the force–dis-
placement relationship:
3EI 3(29, 000, 000 psi)(75 in4 )
Vmax = kuo = 3 uo = 0.1729 in = 193.5 lb
L (15 ft ×12 in/ft )
3
4. Maximum normal stress due to bending.
The maximum bending moment at the top of the column is calculated from equilib-
rium of the column member. As the column has a pin support, summing moments about
the top of the column will result in the following maximum internal bending moment at
the top of the column (see Figure 3.24):
PROBLEMS
3.1 Determine the effective weight at the roof and floor levels for a 60 ft by 150 ft, three-
story office building with equal story heights of 10 ft and the following loading:
Roof DL = 20 psf
Roof LL = 20 psf
Floor DL = 20 psf
Floor LL = 60 psf
Wall DL = 10 psf
Also, the partition load is 10 psf as the occupancy is designated as office building
and the floor live load is <80 psf.
3.2 The bridge beam depicted below supports a rigid deck that weighs 200 kips. Assuming
that the weight is lumped at midspan, determine the natural frequency and period of the
system (E = 29,000 kip/in2 and Ix = 13,000 in4).
50 ft
W = 5 kips
30 ft
3.4 Given the following steel (E = 29,000 kip/in2) building frame, determine the period.
The columns have Ix = 75 in4 and the beam has Ix = 150 in4.
15 kips
15 ft
30 ft
Single-Degree-of-Freedom Structural Dynamic Analysis 87
3.5 The structural system depicted below has a roof weighing 22.5 psf and side sheathing
weighing 10 psf. The total weight is assumed to be concentrated at the bottom of the
roof trusses. The lateral force-resisting system in the North–South direction (one on
each side) consists of bracing with ½-inch steel rods (E = 29,000 ksi). Determine the
natural period of the building.
1 2
4′
12′
A B C
1
N
40′
2 @ 40′
3.6 Given the following structural system and properties, determine the total stiffness and
structural period.
W = 750-kips
3.7 Consider the following building frame with a rigid beam that is pin connected
to one column and rigidly connected to the other as shown; each column has
EI = 40,000,000 kip-in2. Determine the natural frequency and period of the system.
p(t) 200 kips
10 ft pin
15 ft
3.8 After constructing a bridge pier, the owner decides to determine the dynamic proper-
ties of the system. This type of system can be modeled as a concentrated mass atop a
weightless tower. To determine the properties, a cable is attached to the top and pulled
with a crane with a force of 20 kips, which causes a horizontal displacement of 1 in.
The cable then is suddenly cut and the resulting free vibration is recorded. At the end of
10 complete cycles, the time is 2 s and the amplitude is 0.14 in. From this information,
88 Introduction to Earthquake Engineering
compute: (a) undamped natural period, (b) effective stiffness, (c) effective weight, and
(d) effective damping coefficient.
Pier cap
3.9 A wind turbine can be modeled as a concentrated mass atop a weightless tower.
To determine the dynamic properties of the system, a cable is attached to the turbine
and a crane. A lateral force of 200 pounds is applied to the turbine causing a horizontal
displacement of 1.0 in. When the cable is suddenly cut, the resulting free vibration is
recorded. At the end of four complete cycles, the time is 1.25 s and the amplitude is
0.64 in. From this information, compute the: (a) undamped natural period, (b) effective
stiffness, (c) effective weight, and (d) effective damping coefficient.
3.10 To determine the dynamic properties of a simply supported bridge girder, a cable is
attached to its midspan and pulled with a crane. To produce a 4.0-inch deflection at
midspan, the crane exerts a vertical force of 10 kips. When the cable is suddenly cut,
Single-Degree-of-Freedom Structural Dynamic Analysis 89
the resulting free vibration is recorded. At the end of five complete cycles, the time
is 0.25 s and the amplitude is 2.7 in. From this information, compute: (a) undamped
natural period, (b) effective stiffness, (c) effective weight, and (d) effective damping
coefficient.
Instrument
3.13 A 500-pound tiger that was rescued will be placed inside a box and transported on
a flatbed truck. The box will be supported by four springs, each having stiffness of
250 lb/in. The damping for the box supporting system has been estimated at 2%. During
the transport process, it is estimated that the box will experience vertical displace-
ments of u(t) = 1.5 in ⋅ sin 4 rad/s ⋅ t. For the tiger to have a comfortable trip, maximum
relative displacement and velocity of the box should be <0.05 in and 1.0 in/s, respec-
tively. Does the spring damping system satisfy these requirements?
3.14 Given the following building frame with damping of 5% and load p(t) = 200 lbs
sin(5.3 rad/s ⋅ t), determine (a) maximum displacement u, (b) the maximum base shear,
and (c) the maximum normal stresses in the columns (Ix = 75 in4 and Sx = 18 in3 for
W8 × 21). Assume that the beam is rigid.
p(t) 250 kips
pin
10 ft
15 ft
90 Introduction to Earthquake Engineering
3.15 The following frame has a rigid beam, a diagonal 1/2-in diameter steel rod (E = 29,000
ksi) brace, and two steel columns (E = 29,000 ksi and Ix = 82.7 in4). Determine the maxi-
mum force in the diagonal member and the maximum shear force in each of the c olumns
when the frame is subjected to force p(t) = 900 lbs sin 8 rad/s ⋅ t. Damping is 5%.
p(t) 600 kips pin
10 ft
15 ft
pin
20 ft
REFERENCES
American Society of Civil Engineers, ASCE/SEI Minimum Design Loads for Buildings and Other Structures,
Edition, American Society of Civil Engineers, Reston, VA, 2010.
Chopra, A. K., Dynamics of Structures: Theory and Applications to Earthquake Engineering, 4th edition,
Prentice-Hall, Upper Saddle River, NJ, 2012.
Lindeburg, M. R. and K. M. McMullin, Seismic Design of Building Structures: A Professional Introduction
to Earthquake Forces and Design Details, 9th edition, Professional Publications, Inc., Belmont,
CA, 2008.
Silva, P. and S. S. Badie, Optimum beam-to-column stiffness ratio of portal frames under later loads.
Structure Magazine, August 2008.
Villaverde, R., Fundamental Concepts of Earthquake Engineering, CRC Press, Boca Raton, FL, 2009.
4 Response to General Loading
1. Determine the response of a single degree of freedom (SDOF) system to a general load by
the direct and numerical integration of Duhamel’s integral
2. Use shock spectra to solve general dynamics load problems
3. Perform the direct numerical solution of equation of motion to solve for the response of an
SDOF system subjected to a general dynamic load
Thus far, we covered the response of structures to harmonic loadings; however, the ground vibra-
tion of a structure during an earthquake is nonperiodic, and it can be of short or long duration. The
response of an SDOF system subjected to a nonperiodic force can be determined using a convo-
lution integral, Laplace transform, or numerical methods. The convolution integral and Laplace
transform approaches are analytical methods that can be used to evaluate the response of the
system in a closed form when the nonperiodic forcing function can be described analytically. The
analytical approaches are useful in describing the response of an SDOF system and characterizing
the influence of parameters on system response, as well as informing system design. Numerical
methods are often employed in cases where an analytical solution is difficult or not available under
a general forcing function.
In this chapter, we first focus on the use of a convolution integral known as Duhamel’s integral
in the context of structural dynamics to analyze an SDOF system subjected to a general forcing
function. Duhamel’s integral is based on the response of the system due to an impulse, which is a
force that abruptly jumps from zero to a constant value. Several examples using Duhamel’s inte-
gral are outlined in order to demonstrate the analytical solution to nonperiodic, general forcing
functions; these include step, rectangular pulse, and ramp forcing functions. These analyti-
cal examples are used to introduce the concept of the shock spectra for a given forcing function
and how shock spectra can be used to determine the maximum response of structures to such
loadings.
In addition, we demonstrate the use of numerical methods to evaluate the response of an SDOF
system to a seismic ground motion. In particular, we provide examples of how an earthquake
response spectrum is generated for a given earthquake event.
91
92 Introduction to Earthquake Engineering
p(t)
p(t) m
p Impulse = p∆τ
0
τ τ + ∆τ
u(t) p∆τ ⋅ g(t − τ)
Undamped
Damped
0 ⋅ t
u(τ)
du
∑ F = ma : p(t ) = m dt (4.1)
where:
p(t) represents the general loading function
Rearranging Equation 4.1 and integrating the result over the limits of t1 = τ to t2 = τ + Δτ result
in the principle of linear impulse and momentum (the product of mass and velocity), which directly
relates a change in momentum mdu to the impulse applied to the SDOF system p(t)dt, that is, the
magnitude of an impulse is equal to a change in momentum,
Integrating,
t2 τ +Δτ
mdu =
∫ p(t ) dt = ∫ τ
p(t ) dt = pΔτ = impulse (4.3)
t1
p∆τ
du = (4.4)
m
Since the impulse acts for an infinitesimally short duration of time (Δτ is very small), the spring
and damper have no effect during the time the impulse is applied. The impulse essentially imparts a
velocity to the mass of the system. The change in velocity described in Equation 4.4 is equal to the
initial velocity of the system, since prior to time t = τ, the system was at rest.
Considering an underdamped SDOF system, subjected to the impulse at time t = τ as shown in
Figure 4.1, the equation of motion is given by
The initial conditions can now be defined using the initial velocity on the mass at time t = τ,
based on Equation 4.4. The initial velocity is given by
pΔτ
u(τ ) = (4.6)
m
u(τ ) = 0 (4.7)
The solution to the equation of motion is given by the free vibration response of an underdamped
SDOF system shown in Equation 3.36 and rewritten in the following for an impulse applied at time
t = τ :
u(τ ) + ζωnu(τ )
u(t ) = e−ζωn ( t−τ ) sin ω (t − τ ) + u(τ )cos ω (t − τ ) (4.8)
ωD
D D
Substituting the initial conditions from Equations 4.6 and 4.7 into Equation 4.8, the response of
an underdamped SDOF subjected to an impulse at time t = τ is as follows:
pΔτ
u(t ) = e−ζωn ( t−τ ) sin ω (t − τ )
mωD D
(4.9)
Neglecting damping in the system (because the short impulse does not allow the damping
mechanism to absorb much energy), the response function is given by
pΔτ
u(t ) = sin ωn (t − τ ) (4.10)
mωn
The impulse response function of a damped and undamped SDOF system is shown in Figure 4.1.
If a unit impulse is applied to the system, such that pΔτ = 1, then the response of the SDOF
s ystem becomes
1
g(t − τ ) ≡ u(t ) = e−ζωn ( t−τ ) sin ω (t − τ )
mωD D
(4.11)
Here, we define g(t − τ) as the unit impulse response function. The unit impulse response can
be multiplied by the magnitude of the impulse in order to ascertain the response of the system for
damped and undamped cases, as shown in Figure 4.1, u(t) = pΔτ ⋅ g(t − τ).
94 Introduction to Earthquake Engineering
By summing the incremental displacement responses, the total displacement response from all
the incremental impulses can be determined,
u(t ) = ∑ ∆u(t ) ≅ ∑[ p(τ )∆τ ]g(t − τ ) (4.13)
As Δτ becomes infinitesimally small or approaches zero, the summation can be replaced with
integration, yielding the following result:
u(t ) =
∫ p(τ )g(t − τ ) dτ (4.14)
0
Then substituting the definition of the impulse response function from Equation 4.11 into
Equation 4.14, we can determine the total response of an underdamped SDOF system at any time t
as the superposition of the response to all the impulses occurring before t, that is,
p(t)
2
1
t
τ τ + ∆τ
du(t)
Response to impulse at τ
t
τ
u(t)
Total response
t
1
u(t ) =
mωD ∫ p(τ )e −ζωn ( t −τ )
[sin ω D (t − τ )] d τ (4.15)
0
Equation 4.15 is a convolution integral and is known as Duhamel’s integral. It allows for the deter-
mination of the response of an underdamped SDOF system to any general forcing function, and is par-
ticularly useful for determining the effect of random excitations associated with ground motion during
earthquakes. Also, if the forcing function p(t) cannot be expressed by an analytical equation, which is
often the case in earthquake-induced ground motion, the integral of Equation 4.15 can be evaluated
using numerical methods. It is important to recognize that Equation 4.15 assumes that the system is at
rest before the application of the forcing function p(t). Also, Equation 4.15 accounts for the particular
solution of the response up(t) as described in Chapter 3, with the total solution given by Equation 3.23.
If an initial displacement u(0) and initial velocity u( 0) were present, and the complementary (free
vibration response) part of the solution given by Equation 3.36 would be added to Equation 4.15, then
the total response of an underdamped case including initial conditions is
u(0) + ζωnu(0)
u(t ) = e−ζ ωnt sin ω t + u(0)cos ω t
ωD
D D
t
1
+
mωD ∫ p(τ )e −ζωn ( t −τ ) sin ω D (t − τ ) d τ
(4.16)
0
If damping in the system is neglected, then Equations 4.15 and 4.16 can be expressed as follows:
t
1
u(t ) =
mωn ∫ p(τ )[sin ω (t − τ )] dτ
n (4.17)
0
and
t
u(0) 1
u(t ) =
ωn
sin ωnt + u(0)cos ωnt +
mωn ∫ p(τ )[sin ω (t − τ )] dτ
n (4.18)
0
Again, it is important to note that Equation 4.17 assumes that the system is initially at rest,
whereas Equation 4.18 is given to show how the initial displacement, u(0), and the initial velocity,
0), would be included in the response.
u(
EXAMPLE 4.1
Consider the undamped SDOF system, shown in Figure 4.1, subjected to the constant force shown
in Figure E4.1. Assuming that the SDOF system is initially at rest, use Duhamel’s integral to deter-
mine the displacement response, u(t), of the system.
p(t)
po
Solution
1. Describe the forcing function and equation of motion to be solved. The forcing function
p(t) can be described analytically as
p(t ) = po
This forcing function can be considered a step force, or a dynamically applied con-
stant force, because it is abruptly applied to the mass of the system as opposed to a force
that is applied very slowly inducing no vibration on the system and can be considered
statically applied.
In this problem, damping is neglected and so the equation of motion to be solved is
given as
mu(t ) + ku(t ) = po
2. Apply Duhamel’s integral. Since we are neglecting damping effects in the system (ζ = 0)
0) = 0), Equation 4.17 can be applied
and the system is initially at rest (u(0) = 0 and u(
directly. Substituting the forcing function into the equation and integrating results in the
response of the system,
t
1 po t
u(t ) =
mωn ∫ p sin ω (t − τ )dτ = mω
o n 2
n
cos ωn (t − τ )
0
0
or
po
u(t ) = [1− cos ωnt ]
mωn2
po
u(t ) = [1− cos ωnt ] = ust[1− cos ωnt ]
k
where:
ust represents the static displacement of the SDOF system; that is, if the force po were
applied very slowly such that no vibrations were induced, then the displacement
would be ust = po /k
The term [1− cos ωnt ] is the dynamic load factor (DLF) and represents the ampli-
fication of the displacement above the static displacement; that is, u(t)/ust.
3. Plot the response of the SDOF system. By rearranging the response in terms of the
DLF and substituting for ωn = 2π/Tn, the following equation can be used to represent the
response of the SDOF system:
u(t ) t
= DLF = 1− cos 2π
ust Tn
We can use this equation to obtain the displacement response of the SDOF system
subjected to a step force (Figure E4.2).The response of an SDOF system to the step force
p(t) = po is similar to the free vibration response of an undamped system except that it
is shifted up by the static displacement ust. The maximum displacement (umax = 2ust) of
an SDOF system subjected to a suddenly applied force po is twice the displacement of
Response to General Loading 97
1.5
u(t)/ust
0.5
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
t/Tn
the same SDOF system when an equivalent static force po is applied very slowly, and it
occurs at ωt = π or when t = T/2.
EXAMPLE 4.2
Use Duhamel’s integral to determine the response of an undamped SDOF system subjected to a
rectangular impulse that is applied from 0 < t < td. The rectangular impulse is a suddenly applied
constant force, po, which is applied for a specified duration of time, td, as shown in Figure E4.3.
Assume that the SDOF system is initially at rest.
Solution
1. Describe the forcing function and equation of motion to be solved. The forcing function
p(t) can be described analytically as
po t ≤ td
p(t ) =
0 t > td
By superposition, the given forcing function can be described as a sum of step func-
tions where the first forcing function p1(t), with a magnitude of po, starts at t = 0, and the
second forcing function p2(t), with a magnitude of –po, starts at t = td. This is analytically
described as
p1(t ) = po
0 t ≤ td
p2(t ) =
− po t > td
and graphically illustrated using superposition as the sum of step forces shown in
Figure E4.4.
p(t)
po
t
td
po po
= + td
t t t
td
–po
FIGURE E4.4 Illustration of the rectangular impulse as the sum of two step force loading functions.
2. Apply Duhamel’s integral. The response of the SDOF system to the step force p1(t) = po
can be determined by applying Duhamel’s integral and substituting the forcing function.
This result is the same as the response determined in Example 4.1 and rewritten in the
following for convenience:
where:
u1(t) is the response of the SDOF system to the first forcing function p1(t)
The second forcing function p2(t) is a time-delayed step force with a magnitude of
–po, and the response to this second forcing function can also be determined by substi-
tuting the forcing function and evaluating Duhamel’s integral as follows:
t
1 − po t
u2 (t ) =
mωn ∫ − p sin ω (t − τ ) dτ = mω
td
o n 2
n
cos ωn (t − τ )
td
−po
u2(t ) = [1− cos ωn (t − td )] = ust[cos ωn (t − td ) − 1]
k
There are two response regions due to the rectangular pulse: during the pulse when
t ≤ td, the total response of the system is given by u1(t), and after the pulse when t > td, the
total response of the system is given by the two responses, u1(t) + u2(t), that is,
which can be written in terms of the DLF = u(t)/ust and a dimensionless time parameter,
the ratio of duration of the loading to the natural period of the structure, td/Tn:
t
1− cos 2π t ≤ td
Tn
u(t )
DLF = =
ust cos 2π t − td − cos 2π t
t > td
Tn Tn Tn
3. Plot the response of the SDOF system. We can use these equations for different values of
td /Tn to obtain the displacement response of the SDOF system subjected to a rectangular
pulse force as shown in Figure E4.5.
Although we can find closed-form solutions to some general forcing function problems, it is more
efficient to determine the results using computer software, such as MATLAB®. In this chapter, we
use MATLAB to develop short scripts to solve many of the problems encountered in structural
Response to General Loading 99
1.5 2
td td
1 = 0.25 = 0.5
Tn Tn
1
0.5
u(t)/ust
u(t)/ust
0 0
–0.5
–1
–1
–1.5 –2
0 1 2 3 4 5 0 1 2 3 4 5
t/Tn t/Tn
2 2
td td
=1 =2
1.5 Tn 1.5 Tn
1 1
u(t)/ust
0 0
–0.5 –0.5
0 1 2 3 4 5 0 1 2 3 4 5
t/Tn t/Tn
FIGURE E4.5 Response of an SDOF system due to a rectangular pulse and varying pulse durations.
dynamics. And in later chapters, we will use these MATLAB scripts to solve earthquake engineer-
ing problems. These programs have powerful algorithms that can perform complex calculations
efficiently. In fact, modern analysis and design have become highly dependent on these computa-
tional tools.
MATLAB (MATrix LABoratory) was originally designed as an interactive software system
for matrix operations, such as solving systems of linear equations, and computing eigenvalues and
eigenvectors. More recent versions of the program include extensive graphics capabilities that can
be used to easily plot results, such as those shown in Examples 4.1 and 4.2. It also has an extensive
library of functions, including the one that performs convolution integrals numerically (the conv
function), as well as differentiation and integration. To facilitate the development of a script to
perform the convolution operation, we first rewrite Equation 4.17 (or Equation 4.15) in terms of two
functions that are then discretized into dimensionless vectors. The first function includes the shape
of the pulse and is written as a generalized forcing function p(t) = pox(t) that can be substituted into
Equation 4.17 (as p(τ) = pox(τ)),
t
p
u(t ) = o
mωn ∫ x(τ ) ⋅ sin ω (t − τ ) dτ
n (4.19)
0
We also write this equation in terms of a dimensionless time factor, the ratio of time to the natu-
ral period of the structure, t/Tn. That is, substituting ωn = 2π/Tn and ωn = k /m , we obtain
t t
ω p 2π po 2π
u(t ) = n o
ωn mωn ∫
0
x(τ ) ⋅ sin ωn (t − τ ) d τ =
Tn m
( k /m ) ∫
x(τ ) ⋅ sin
2
0
T n
(t − τ ) d τ
100 Introduction to Earthquake Engineering
or
t
u(t ) τ t τ
ust
= 2π
∫ x ⋅ sin 2π − d τ
Tn Tn Tn
(4.20)
0
where ust = po/k. Equation 4.20 can be written in convolution form as
t
u(t ) τ t τ
DLF =
ust
= 2π
∫ x ⋅ h 2π − dτ
Tn Tn Tn
(4.21)
0
where the function h(ξ) = sin(ξ), and ξ is the argument of the function, which is the quantity in the
brackets in this case.
EXAMPLE 4.3
Use the convolution function in MATLAB and Equation 4.21 to solve for the response due to the
rectangular impulse shown in Figure E4.3.
Solution
1 t ≤ td
x(t ) =
0 t > td
1. The MATLAB script is as follows:
clear all
clc
% specify the duration of the pulse td/T
tdT = input ('Enter the value for td/T ratio: ');
% create a vector with n equally spaced points to represent t/Tn
n = 500;
tT = linspace(0,5,n);% loop over t/Tn to create vector of input pulse
for i = 1:length(tT)
if tT(i) <= tdT
p(i) = 1;
else
p(i) = 0;
end
end
% integrate pulse function to get response
dt=tT(2)-tT(1);
p=p*dt;
h=sin(2*pi*tT);
uust=2*pi*conv(p,h);
%create plot
plot (tT, uust(1:length(tT)), 'LineWidth',2, 'Color',[0 0 0]);
xlabel ('t/T_n', 'FontSize', 12, 'FontName', 'Times New Roman',…
'FontAngle', 'italic');
ylabel ('u(t)/u_{st}', 'FontSize', 12, 'FontName', 'Times New Roman', …
'FontAngle', 'italic');
This script can be used to generate graphs of the normalized displacement response
of SDOF systems subjected to a rectangular impulse, identical to those obtained using
the results of Example 2.
EXAMPLE 4.4
Use the convolution function of MATLAB and Equation 4.21 to determine graphically the dynamic
response of a tower subjected to the load shown in Figure E4.6.
Response to General Loading 101
p(t)
p(t)
W = 38.6 kips
120k
k = 100 kips/in
t(s)
0.02 0.04 0.06
Solution
1. Mass, stiffness, and natural frequency of the SDOF system. The tower can be modeled
as an SDOF system. The stiffness of the system is given as 100 kips/in and the mass can
be calculated as follows:
The natural frequency of the SDOF system is determined using Equation 3.2,
k 100, 000
ωn = = = 31.6 rad/s
m 100
50t t ≤ 0.02 s
0.02 s < t ≤ 0.04 s
1
x(t ) =
3 − 50t 0.04 s < t ≤ 0.06 s
0 t > 0.06 s
Also, po = 120 kips
3. Displacement response as a function of time script:
clear all
clc
% input data
omega = 31.6; % natural frequency in rad/sec
po=120000; % peak force in lbs
m=100; % mass in lbs per square second
dt=0.002; % time increment for calculations
% loop over t to create vector based on pulse function values
for i = 1:500;
t(i)=dt*i;
if t(i) <= 0.02
p(i) = 50*t(i);
elseif t(i)<= 0.04
p(i) = 1;
elseif t(i)<= 0.06
p(i) = 3-50*t(i);
else
p(i) = 0;
end
end
102 Introduction to Earthquake Engineering
u(in)
0
–1
–2
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
t(s)
p=p*dt;
h=sin(omega*t);
uust=po*conv(p,h)/(m*omega);
uustr = max(abs(uust)) % maximum displacement
%create plot
plot (t, uust(1:500), 'LineWidth',2, 'Color',[0 0 0]);
xlabel ('t(sec)', 'FontSize', 12, 'FontName', 'Times New Roman',...
'FontAngle', 'italic');
ylabel ('u(in)', 'FontSize', 12, 'FontName', 'Times New Roman', ...
'FontAngle', 'italic');
EXAMPLE 4.5
Use the results from Example 4.2 to draw the shock spectrum for the rectangular pulse force, as
shown in Figure E4.2.
Response to General Loading 103
Solution
1. Use the DLF versus td/Tn results obtained in part 2 of Solution 4.2; that is,
t
1− cos 2π t ≤ td
u(t )
Tn
DLF = =
ust t t t
cos 2π − d − cos 2π t > td
Tn Tn Tn
The maximum values of the DLF, DLFmax, for various values of td/Tn can be obtained
more efficiently using MATLAB. The results of DLFmax versus td/Tn can then be graphed
to obtain the shock spectrum of a pulse loading, as shown in the Figure E4.8. The script
is as follows:
clear all
clc
% create two vectors of equal n length to represent td/Tn and t/Tn
n = 500;
tdT = linspace(0,5,n);
tT = linspace(0,5,n);
% loop over td/Tn and t/Tn to create the shock spectrum
for j = 1:n
for i = 1:n
if tT(i) <= tdT(j)
uust(i) = 1 - cos(2*pi*tT(i));
else
uust(i) = cos(2*pi*(tT(i)-tdT(j)))- cos(2*pi*tT(i));
end
end
uustr(j)=max(abs(uust)); % Select the max values from each response
end
% create figure
figure1=figure;
axes1 = axes('Parent',figure1);
xlim (axes1,[0 5]);
ylim (axes1,[-2.5 2.5]);
box (axes1, 'on');
p(t)
1.5
po
DLFmax
1
td t
0.5
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
td/Tn
FIGURE E4.8 Shock spectrum of an SDOF system subject to a rectangular impulse load.
104 Introduction to Earthquake Engineering
EXAMPLE 4.6
Write a MATLAB script using the conv function to plot the shock spectrum for the rectangular
pulse in Example 4.2.
Solution
clear all
clc
% create two vectors of equal n length to represent td/Tn and t/Tn
n = 500;
tdT = linspace(0,5,n);
tT = linspace(0,5,n);
% loop over td/Tn and t/Tn to create the shock spectrum
for j = 1:n
for i = 1:n
if tT(i) <= tdT(j)
p(i) = 1;
else
p(i) = 0;
end
end
% integrate pulse function to get response
dt=tT(2)-tT(1);
p=p*dt;
h=sin(2*pi*tT);
uust=2*pi*conv(p,h);
uustr(j)=max(abs(uust)); % Select the max values from each response
end
% create figure
figure1=figure;
axes1 = axes('Parent',figure1);
xlim (axes1,[0 5]);
ylim (axes1,[0 2.2]);
box (axes1, 'on');
grid (axes1, 'on');
hold (axes1, 'all');
%create plot of the shock spectrum
plot (tdT,uustr,'LineWidth',2,'Color',[0 0 0]);
xlabel ('t_d /T_n','FontSize',12,'FontName','Times New Roman',...
'FontAngle','italic');
ylabel ('DLF_{max}','FontSize',12,'FontName','Times New Roman',...
'FontAngle','italic');
This script creates a graph identical to the graph shown in Figure E4.8. The difference in the
two scripts is in the way they establish the two arrays that are plotted to generate the figure.
Example 4.5 uses the analytical solution, while Example 4.6 applies the conv function to obtain
the response array.
EXAMPLE 4.7
Draw the shock spectrum for the response of an SDOF system subjected to the symmetric trian-
gular pulse loading function shown in Figure E4.9.
Response to General Loading 105
p(t)
po
t
td/2 td
Solution
1. Obtain function x(t):
2t /td t ≤ td / 2
x(t ) = 2 − 2t /td td / 2 < t ≤ td
0 t > td
clear all
clc
% create two vectors of equal n length to represent td/Tn and t/Tn
n = 500;
tdT = linspace(0,5,n);
tT = linspace(0,5,n);
% loop over td/Tn and t/Tn to create the shock spectrum
for j = 1:n
for i = 1:n
if tT(i) <= tdT(j)/2
p(i) = 2*tT(i)/tdT(j);
elseif tT(i) <= tdT(j)
p(i) = 2-2*tT(i)/tdT(j);
1.6
1.4
1.2
1
DLFmax
p(t)
0.8
po
0.6
0.4
0.2
td/2 td t
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
td /Tn
FIGURE E4.10 Shock spectrum of an SDOF system subject to a symmetric triangular impulse load.
106 Introduction to Earthquake Engineering
else
p(i) = 0;
end
end
% integrate pulse function to get response
dt=tT(2)-tT(1);
p=p*dt;
h=sin(2*pi*tT);
uust=2*pi*conv(p,h);
uustr(j)=max(abs(uust)); % Select the max values from each response
end
% create figure
figure1=figure;
axes1 = axes('Parent',figure1);
xlim (axes1,[0 5]);
ylim (axes1);
box (axes1, 'on');
grid (axes1, 'on');
hold (axes1, 'all');
%create plot of the shock spectrum
plot (tdT,uustr,'LineWidth',2,'Color',[0 0 0]);
xlabel ('t_d /T_n','FontSize',12,'FontName','Times New Roman',...
'FontAngle','italic');
ylabel ('DLF_{max}','FontSize',12,'FontName','Times New Roman',...
'FontAngle','italic');
EXAMPLE 4.8
Draw the shock spectrum for the response of an SDOF system subjected to the triangular pulse
loading function shown in Figure E4.11.
Solution
1. Obtain function x(t):
1− t /td t ≤ td
x(t ) =
0 t > td
clear all
clc
% create two vectors of equal n length to represent td/Tn and t/Tn
n = 500;
p(t)
po
t
td
1.8
1.6
1.4
DLFmax 1.2
p(t)
1
po
0.8
0.6
0.4
td t
0.2
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
td /Tn
FIGURE E4.12 Shock spectrum of an SDOF system subject to a triangular impulse load.
tdT = linspace(0,5,n);
tT = linspace(0,5,n);
% loop over td/Tn and t/Tn to create the shock spectrum
for j = 1:n
for i = 1:n
if tT(i) <= tdT(j)
p(i) = 1-tT(i)/tdT(j);
else
p(i) = 0;
end
end
% integrate pulse function to get response
dt=tT(2)-tT(1);
p=p*dt;
h=sin(2*pi*tT);
uust=2*pi*conv(p,h);
uustr(j)=max(abs(uust)); % Select the max values from each response
end
% create figure
figure1=figure;
axes1 = axes('Parent',figure1);
xlim (axes1,[0 5]);
ylim (axes1);
box (axes1, 'on');
grid (axes1, 'on');
hold (axes1, 'all');
%create plot of the shock spectrum
plot (tdT,uustr,'LineWidth',2,'Color',[0 0 0]);
xlabel ('t_d /T_n','FontSize',12,'FontName','Times New Roman',…
'FontAngle','italic');
ylabel ('DLF_{max}','FontSize',12,'FontName','Times New Roman',…
'FontAngle','italic');
EXAMPLE 4.9
Draw the shock spectrum for the response of an SDOF system subjected to the half-cycle sine
pulse force shown in Figure E4.13.
108 Introduction to Earthquake Engineering
p(t)
po
t
td
Solution
1. Obtain function x(t):
sin(πt /td ) t ≤ td
x(t ) =
0 t > td
clear all
clc
% create two vectors of equal n length to represent td/Tn and t/Tn
n = 500;
tdT = linspace(0,5,n);
tT = linspace(0,5,n);
% loop over td/Tn and t/Tn to create the shock spectrum
for j = 1:n
for i = 1:n
if tT(i) <= tdT(j)
p(i) = sin(pi*tT(i)/tdT(j));
else
p(i) = 0;
end
end
% integrate pulse function to get response
dt=tT(2)-tT(1);
1.8
1.6
1.4
1.2
1
DLFmax
0.8
p(t)
0.6 po
0.4
0.2
td t
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
td/Tn
FIGURE E4.14 Shock spectrum of an SDOF system subject to a half-cycle sine impulse load.
Response to General Loading 109
p=p*dt;
h=sin(2*pi*tT);
uust=2*pi*conv(p,h);
uustr(j)=max(abs(uust)); % Select the max values from each response
end
% create figure
figure1=figure;
axes1 = axes('Parent',figure1);
xlim (axes1,[0 5]);
ylim (axes1);
box (axes1, 'on');
grid (axes1, 'on');
hold (axes1, 'all');
%create plot of the shock spectrum
plot (tdT,uustr,'LineWidth',2,'Color',[0 0 0]);
xlabel ('t_d /T_n','FontSize',12,'FontName','Times New Roman',...
'FontAngle','italic');
ylabel ('DLF_{max}','FontSize',12,'FontName','Times New Roman',...
'FontAngle','italic')
EXAMPLE 4.10
Given the building frame shown in Figure E4.15, which is subjected to a triangular impulse force
(see Figure E4.11) of amplitude po = 5 kips and duration td = 0.6 s, determine (1) maximum dis-
placement at the top, (2) the maximum base shear, and (3) the maximum bending stresses in the
columns (Ix = 82.7 in4 and Sx = 20.9 in3 for W8 × 24). Assume that the beam is rigid.
Solution
1. Mass, stiffness, and natural period and frequency of the SDOF system. The building
frame can be modeled as an SDOF system, assuming that only lateral deformations of
the columns occur. The stiffness of the system is the sum of column lateral stiffnesses.
The mass and stiffness of the SDOF system are calculated as follows:
Mass:
12EI 3EI 12( 29, 000, 000 psi)(82.7 in4 ) 3( 29, 000, 000 psi)(82.7 in4 )
k= + 3 = + = 5455 lb/in
h13 h2 (15 ft × 12 in/ft )3 (20 ft × 12 in/ft )3
u
20 kips
F(t)
15 ft W8 × 24
20 ft
The natural frequency of the SDOF system is determined using Equation 3.2,
k 5455
ωn = = = 10.3 rad/s
m 51.8
The natural period of the SDOF system is determined using Equation 3.5,
2π 2π
Tn = = = 0.61 s
ωn 10.3 rad/s
2. Determine the maximum displacement. The maximum displacement due to the applied
triangular pulse forcing function can be determined using the shock spectrum obtained
in Figure E4.12. This figure provides the maximum dynamic load factor, DLFmax = uo /ust
as a function of the ratio of pulse duration to natural period, td /Tn, where the equivalent
static displacement ust can be calculated using the basic force−displacement relationship
for the SDOF system as follows:
Also, the ratio of pulse duration to natural period can be determined as follows,
td 0 .6
= = 0.98
Tn 0.61 s
3. Maximum base shear. Calculate the maximum base shear at the support for each column
using the force−displacement relationship or the equilibrium of the column member as
described in Figure 3.24:
The 15-foot column:
4. Maximum normal stress due to bending. The maximum bending moment at the top of
each column is calculated from the equilibrium of the column member as described in
Figure 3.24.
The 15-foot column:
The maximum moment yields the maximum bending stress; thus, the 15-foot column
experiences the maximum stress. Applying the flexure formula from mechanics of mate-
rials (Equation 3.60), the normal stress in this column due to the bending moment is
M1max 631kip ⋅ in
σ max = = = 30.2 ksi
S 20.9 in3
The total response of an underdamped SDOF system including initial conditions subjected to a
ground excitation, ug (t ), can be determined using Equation 4.16 with a new forcing function
When the system starts from rest, we can determine the response due to a ground excitation for
an underdamped system in terms of the relative displacement of the mass with respect to the ground
by substituting Equation 4.22 into Duhamel’s integral (Equation 4.15). That is,
t
1
u(t ) = −
ωD ∫ u (t )e
g
−ζ ωn ( t −τ )
[sin ωD (t − τ )] d τ (4.24)
0
EXAMPLE 4.11
Draw the response of the frame shown in Figure E4.15, when subjected to the ground motion of
the North−South component of the horizontal ground acceleration recorded at the Imperial Valley
Irrigation District substation, El Centro, California, during the Imperial Valley Earthquake of May
18, 1940, from here on referred to as the El Centro earthquake. Assume 2% damping. The input
ground acceleration was obtained from http://www.vibrationdata.com/elcentro.dat, and is graphi-
cally shown in Figure E4.16.
Solution
1. The natural frequency of the frame ωn was obtained in Example 4.10 as 10.3 rad/s.
Using the conv function introduced in Section 4.2, we can write the following script to
perform the analysis:
clear all
clc
dam = 0.02; % given damping of 2%
wn = 10.3; % rad/sec, determined in Example 10
load elcentro.mat; % load El Centro data
N = length(elcentro); % number of points in the ground acceleration file
112 Introduction to Earthquake Engineering
üg(g)
0.4
0.3
0.2
0.1
0
–0.1 0 5 10 15 20 25 30
–0.2
–0.3
t(s)
–0.4
FIGURE E4.16 Ground motion acceleration time history for El Centro earthquake.
This gives a maximum displacement of 3.12 in, and the graphical response history
results are shown in Figure E4.17.
2
u(in)
–2
–3.12 in
–4
0 2 4 6 8 10 12 14 16 18 20
t(s)
FIGURE E4.17 Displacement response of the SDOF building frame due to El Centro earthquake.
Response to General Loading 113
Equation 4.23 can be rewritten as the DLF in terms of the relative displacement uo/ust or
the total accelerations uot /ugo . Additionally, we can normalize time with respect to the natural
period as a dimensionless time parameter, t/Tn. First, let ust = peff/k; assuming peff = mugo leads to
ust = mugo /k = ugo /ωn2, since k = ωn2 m and ugo is the peak value of the ground acceleration time his-
tory, ug . The DLF can then be written as
u ω 2u ut
DLF = o = n o = o (4.25)
ust ugo ugo
where uot is the total acceleration, which is related to the relative displacement by ωn2uo. From
Equation 4.21, we determine the convolution integral component of Equation 4.24,
t
τ t τ
DLF = 2 π
∫ x ⋅ h 2π − d τ
Tn Tn Tn
(4.26)
0
where x(ξ) is a function that describes the applied displacement (or acceleration) history function.
This implies that all of the shock spectra derived thus far are applicable to ground excitation forcing
functions as well.
However, generating a shock spectrum using the conv function in MATLAB is relatively simple.
For instance, in the following example, we use the El Centro earthquake ground acceleration pre-
sented in Example 4.11. This data will be used extensively in the remainder of the book. Also, the
terminology changes when dealing with seismic response; rather than shock spectrum, it is custom-
ary (and more appropriate) to refer to the shock spectrum as the response spectrum. The response
spectrum is the plot of the maximum response as a function of the SDOF system period. This topic
will be covered extensively in Chapter 5, including the development of the design response spec-
trum for elastic and inelastic structural systems.
EXAMPLE 4.12
Draw the response spectrum for the response of an SDOF system subjected to the El Centro
earthquake ground acceleration.
Solution
The input ground acceleration can be obtained from http://www.vibrationdata.com/elcentro.dat
as a fraction of the acceleration due to gravity, g. In this case, we derive the response based on
Equation 4.23 and the natural period since we do not have a lapse time for the input forcing
function. Thus, the displacement response as a function of the natural period script is:
clear all
clc
dam = 0.02; % damping; can changed to generate different spectra
load elcentro.mat; % load El Centro data
N = length(elcentro); % number of points in the ground acceleration file
acc = elcentro; % ground acceleration data
DT = 0.02; % sampling rate
for i=N:-1:1
t(i)=i*DT;
end
for j = 500:-1:1
period(j)=j*DT/2; % natural period array
wn(j)=2*pi/period(j); % natural frequency array
p=acc*DT;
h=exp(-dam*wn(j)*t).*sin(wn(j)*t);
u=conv(p,h);
uustr(j)=max(abs(u))/wn(j); % maximum displacement
114 Introduction to Earthquake Engineering
end
%create plot
plot (period, uustr*386.4, 'LineWidth',2, 'Color',[0 0 0]);
xlabel ('T_n(sec)','FontName','Times New Roman','FontAngle','italic');
ylabel ('u(in)','FontName','Times New Roman','FontAngle','italic');
∆a
a(τ ) = ai + i τ , ti ≤ τ ≤ ti+1 (4.27)
∆t
a(t) ai+1
ai
a(τ )
∆ ti
ti ti+1 t
τ
where:
∆ai = ai +1 − ai
ti = i ⋅ ∆t
and
∆t = ti +1 − ti
for i = 1, 2, 3, … ,N
∆a
u + 2ωnζυ + ωn2u = −ai − i τ , ti ≤ τ ≤ ti+1 (4.28)
∆t
The solution to this linear differential equation with constant coefficients is given by the
complementary and particular components,
u(t ) = uc (t ) + u p (t ) (4.29)
u p (τ ) = Bi + Ai (τ − ti )
Here, Ai and Bi are the constants of integration, determined by substituting up back into the equa-
tion of motion. After applying boundary conditions, the formulas to calculate the displacement,
velocity, and acceleration at time step ti+1 = ti + Δt are
respectively. Where the coefficients of the displacements and velocities in the first two equations
only need to be computed once and are given as
ζ
a11 = e−ζωn∆t sin ωD∆t + cos ωD∆t
1 − ζ 2
sin ωD∆t
a12 = e−ζ ωn∆t
ωD
116 Introduction to Earthquake Engineering
ωn
a21 = − e−ζωn∆t sin ωD∆t
2
1− ζ
ζ
a22 = e−ζωn∆t cos ωD∆t − sin ωD∆t
1− ζ 2
2ζ 2 −1 ζ sin ω ∆t 2ζ 1 2ζ
b11 = e−ζ ωn∆t 2 + D
− 3 + 2 cosωD∆t − 3
ωn ∆t ωn ω D ωn∆t ωn ωn∆t
2ζ 2 −1 sin ω ∆t 2ζ 1 2ζ
b12 = −e−ζωn∆t 2 D
+ 3 cos ωD∆t − 2 + 3
ωn ∆t ωD ωn∆t ωn ωn∆t
2
2ζ −1 ζ ζ
b21 = e−ζωn ∆t 2 + cos ωD∆t − sin ωD∆t
ωn ∆t ω n 1− ζ 2
2ζ 1 1
− 3 + 2 (ωD sin ωD∆t + ζ ωn cos ωD∆t ) + 2
ωn∆t ωn ωn ∆t
e−ζωn∆t ζ − 1
b22 = sin ω D ∆t + cos ω D ∆t ω 2∆t
2
ωn ∆t 1 − ζ 2
n
EXAMPLE 4.13
Write a MATLAB script for the Nigam−Jennings algorithm; then use the script to draw displace-
ment, velocity, and acceleration response spectra for the response of the SDOF system subjected
to the El Centro earthquake ground acceleration presented in Example 4.11.
Solution
Again, the input ground acceleration can be obtained from http://www.vibrationdata.com/
elcentro.dat as a fraction of the acceleration due to gravity g. In this example, we write a MATLAB
script based on Equations 4.30 and 4.31. The maximum displacement, velocity, and acceleration
response as a function of the natural period script is:
clear all
clc
d(1) = 0;% initial displacement
v(1) = 0;% initial velocity
dam = 0.02; % damping; can be changed to generate different response spectra
DT = 0.02; % sampling rate
load elcentro.mat; % load El Centro data
N = length(elcentro); % number of points in the ground acceleration file
acc = elcentro; % ground acceleration data
% generate the data to graph a response spectrum by changing Omega
for j = 250:-1:1
period(j) = j*DT; % natural period
wd = 2*pi/period(j)*sqrt(1-dam^2); % damped frequency
wn = 2*pi/period(j); % natural frequency
Response to General Loading 117
%%% compute values of the response and choose the maximum %%%
% first calculate the elements of the matrices A and B
a11=exp(−dam*wn*DT)*(dam/sqrt(1-dam^2)*sin(wd*DT)+cos(wd*DT));
a12=exp(−dam*wn*DT)/wd*sin(wd*DT);
a21=-wn/sqrt(1-dam^2)*exp(-dam*wn*DT)*sin(wd*DT);
a22=exp(−dam*wn*DT)*(cos(wd*DT)−dam/sqrt(1−dam^2)*sin(wd*DT));
b11=exp(−dam*wn*DT)*(((2*dam^2−1)/(wn^2*DT)+dam/wn)*sin(wd*DT)/wd+...
(2*dam/(wn^3*DT)+1/wn^2)*cos(wd*DT))−2*dam/(wn^3*DT);
b12=-exp(−dam*wn*DT)*((2*dam^2−1)/(wn^2*DT)*sin(wd*DT)/wd+...
2*dam/(wn^3*DT)*cos(wd*DT))−1/wn^2+2*dam/(wn^3*DT);
b21=exp(−dam*wn*DT)*(((2*dam^2−1)/(wn^2*DT)+dam/wn)*(cos(wd*DT)-...
dam/sqrt(1−dam^2)*sin(wd*DT))−(2*dam/(wn^3*DT)+1/wn^2)*(wd*sin(wd*DT)+...
dam*wn*cos(wd*DT)))+1/(wn^2*DT);
b22=(−1+exp(-dam*wn*DT)*(dam/sqrt(1-dam^2)*sin(wd*DT)+cos(wd*DT)))/(wn^2*DT);
%loop to find response
for i = 1:N-1
d(i+1) = a11*d(i)+a12*v(i)+b11*acc(i)+b12*acc(i+1);
v(i+1) = a21*d(i)+a22*v(i)+b21*acc(i)+b22*acc(i+1);
Responseaccelation(i) = -2*wn*dam*v(i+1)-(wn^2)*d(i+1);
end
% compute the value of the largest response
max_acc(j) = max(abs(Responseaccelation));
max_vel(j) = max(abs(v));
max_dis(j) = max(abs(d));
end
subplot(3,1,1), plot(period,max_dis*386.4)
ylabel('Displacement(in)','FontName','Times New Roman','FontAngle','italic')
subplot(3,1,2), plot(period,max_vel*386.4)
ylabel('Velocity(in/s)','FontName','Times New Roman','FontAngle','italic')
subplot(3,1,3), plot(period,max_acc)
xlabel('T_n(sec)','FontName','Times New Roman','FontAngle', 'italic')
ylabel('Acceleration(g)','FontName','Times New Roman','FontAngle','italic')
The results for the displacement response spectrum obtained using this method and the conv
f unction in MATLAB, shown in Figure E4.18, are virtually identical. The other spectrum are shown
in Figure E4.19.
20
15
u(in)
10
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Tn(s)
FIGURE E4.18 Displacement response spectrum for an SDOF system due to El Centro earthquake.
118 Introduction to Earthquake Engineering
20
Displacement (in)
15
10
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
50
40
Velocity (in/s)
30
20
10
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
1.5
Acceleration (g)
0.5
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Tn(s)
FIGURE E4.19 Displacement, velocity, and acceleration response spectra for the SDOF system
subjected to El Centro earthquake.
and acceleration as shown in this figure into the equation of motion, Equation 3.49, we obtain the
equation of motion for the time interval 2Δt,
where ugi is the ground acceleration at time step i. Also, ui and ui –1 are known from the preceding
step. Rearranging Equation 4.33 to have these known quantities on the right,
1 ω ζ 1 ω ζ 2
+ n u = −ugi − − n u − ω 2 − u (4.34)
(Δt )2 Δt i+1 (Δt )2 Δt i−1 n (Δt )2 i
u(t) ui+1
ui+1 – ui–1
u⋅ i =
2∆t
ui–1
u – 2ui + ui–1 ∆t ∆t
u⋅⋅i = i+1
(∆t)2
t
ti – ∆t ti ti + ∆t
or
ˆ
kui +1 = a
ˆi (4.35)
where:
1 ωζ
kˆ = 2
+ n
(Δt ) Δt
1 ω ζ 2
aˆi = −ugi − − n ui−1 − ωn2 − ui = −ugi − aui−1 − bui
(Δt ) 2
Δt (Δt )2
Determining the initial velocity uo and displacement uo from the specified initial conditions,
we can use the following algorithm for the central difference method:
(∆t )2
u−1 = uo − ∆t ⋅ uo + ⋅ uo
2
1 2ω ζ
kˆ = 2
+ n
(Δt ) 2Δt
1 ωζ
a= 2
− n
(Δt ) Δt
2
b = ω 2n−
(Δt )2
aˆi
ui +1 =
kˆ
ui +1 − ui −1
ui =
2∆t
ui +1 − 2ui + ui −1
ui =
(∆t )2
This method is conditionally stable. It requires the following time step length for stability.
Tn (4.36)
Δt <
π
120 Introduction to Earthquake Engineering
This is rarely a limitation in seismic response analyses since Δt is typically specified as 0.02 s or
smaller to accurately define the ground acceleration ug .
Where parameters β and γ define the variation of the acceleration over nonconstant time inter-
vals Δti. The values of these parameters must be chosen carefully in order to obtain stable results.
Two sets of values that produce stable results include β = 1/2 and γ = 1/4 (constant average accelera-
tion) and β = ½ and γ = 1/6 (linear acceleration).
After obtaining the initial velocity uo and displacement uo from the specified initial conditions,
we can use the following algorithm for the Newmark Beta method:
1. Initial calculations based on the known quantities and time intervals Δti:
1 2ω ζγ
a1 = + n
β (Δt )2 β ⋅ Δt
1 γ
a2 = + 2ωnζ −1
β ⋅ Δt
β
1 γ
a3 = −1 + 2ωnζ −1Δt
2β
2β
k̂ = ω 2n+ a1
aˆi +1
ui +1 =
kˆ
γ γ γ
ui +1 = (ui +1 − ui ) + 1 − ui + Δt 1 − ui
β ⋅ Δt
β
2β
Response to General Loading 121
(ui +1 − ui ) u 1
ui +1 = − i − −1 ui
β (Δt ) 2
β ⋅ Δt 2β
This method is also conditionally stable. It requires the following time step for stability:
Δt ≤ Tn π 2(γ − 2β ) (4.39)
PROBLEMS
4.1 Use Duhamel’s integral to determine the response of an undamped SDOF system
subjected to a half-sine impulse that is applied from 0 < t < t d. Assume that the SDOF
system is initially at rest and t d /T ≠ 0.5. Hint, this can be solved by using superposition
of two sinusoidal excitations.
4.2 Use Duhamel’s integral to determine the dynamic response of a tower subjected to the
load shown.
p(t) p(t)
W = 38.6 kips
120
k = 100 kips/in
0.04 t(s)
0.02
4.3 Draw the shock spectrum for an SDOF system for the triangular impulse loading shown.
p(t)
po
td t
4.4 Assume that Burn’s tower (a water tank weighing 4200 kips) on the campus of University
of the Pacific is supported on a 120 ft tall cantilever tower with stiffness of 2000 kips/in.,
determine the design values for lateral deformation and base shear for a symmetric tri-
angular impulse of amplitude 100 kips and duration t d = 0.1 s.
4.5 The bridge beam depicted below supports a rigid deck that weighs 200 kips. The beam
is subjected to the blast load described by the triangular impulse loading shown.
Determine the maximum stress in the beam. (E = 29,000 kip/in2, Ix = 13,000 in4, and
Sx = 830 in3).
122 Introduction to Earthquake Engineering
p(t)
50 ft
p(t)
100 kips
t (s)
0.5
4.6 Consider the following building frame with a rigid beam that is pinned con-
nected to one column and rigidly connected to the other as shown; each column has
EI = 40,000,000 kip in2. The frame is subjected to the blast load shown at the beam
level. Determine the maximum stress in each column given Sx = 2000 in3.
10 ft Pin
1000 kips
15 ft
t(s)
0.13
4.7 Write a MATLAB script for the central difference algorithm; then use the script to
draw displacement, velocity, and acceleration response spectra for the response of the
SDOF system subjected to the El Centro earthquake ground acceleration presented in
Example 4.11.
4.8 Write a MATLAB script for the Newmark Beta algorithm; then use the script to
draw displacement, velocity, and acceleration response spectra for the response of the
SDOF system subjected to the El Centro earthquake ground acceleration presented in
Example 4.11.
4.9 Write a MATLAB script to solve Problem 4.8 using the lsim function in MATLAB,
which simulates time response of dynamic systems to arbitrary inputs.
REFERENCE
Chopra, A. K., Dynamics of Structures: Theory and Applications to Earthquake Engineering, 4th edition,
Prentice-Hall, Upper Saddle River, NJ, 2012.
5 Response Spectrum
Analysis of SDOF System
After reading this chapter, you will be able to:
1. Use response spectra to calculate the maximum response of a structure, specifically dis-
placements and base shear
2. Obtain a design elastic response spectrum
3. Use design response spectra to calculate maximum response of a structure, specifically
displacements and base shear
4. Develop an inelastic response spectrum
5. Use an inelastic response spectrum to calculate the maximum response of a structure,
specifically displacements and base shear
Given the complexity of seismic ground acceleration records, numerical methods such as those
introduced in Chapter 4 must be applied to obtain the response of a structure to an input ground
excitation (see Figure 5.1). The response depicted in Figure 5.1 is analogous to the analysis per-
formed in Example 4.11 of Chapter 4. For seismic-resistant design, the entire response history of
the system is not necessary, only the absolute maximum value of the response is required. However,
it is necessary to process a large portion of the response history to identify this maximum response
value. Without the aid of computer software, this process can be time consuming, even for the
simple impulsive loadings presented in Chapter 4. Also, the response (displacement, velocity, or
acceleration) of a structure to a seismic excitation depends on system dynamic properties (mass,
stiffness, and damping) and ground excitation characteristics. As discussed in Chapter 4, it is more
convenient to use a plot of the maximum response to a specified loading for all SDOF systems
based on their inherent natural period (or frequency) and damping; the so-called shock spectrum
for impulse loading and response spectrum for seismic loading. Here, we review the theoretical con-
cepts as well as a practical approach for developing a response spectrum that can be used for design;
this is the precursor to the contemporary design response spectrum used in current design codes
(see Chapter 8). Since most systems are allowed to deform inelastically during a seismic event, we
also present details of the design response spectrum for systems in the inelastic range.
123
124 Introduction to Earthquake Engineering
ut
Response action
Tn = 1.0 s and ζ = 0.05 u u (in)
5 Max value
m
–5
ug(t)
–0.5
....
u umaxD
Tn = 3 s t D
umax = Sd = D
D
A
Tn (s)
0.2 0.5 1 3
Vb(t) Mb(t)
FIGURE 5.4 Equivalent static force, base shear, and overturning moment.
EXAMPLE 5.1
Use the MATLAB® script provided in Chapter 4, Example 4.12 to draw response spectrum of the
displacement, velocity, and acceleration for an SDOF system subjected to the El Centro earth-
quake ground acceleration for 2%, 5%, and 10% damping.
Solution
Input ground acceleration is given in Figure E5.1.
200
üg (in/s/s)
–200
0 5 10 15 20 25 30
t (s)
20
ζ = 0.02
15
ζ = 0.05
Sd (in)
10 ζ = 0.1
0
0 0.5 1 1.5 2 2.5 3
60
40
Sv (in/s)
20
0
0 0.5 1 1.5 2 2.5 3
1.5
1
Sa (g)
0.5
0
0 0.5 1 1.5 2 2.5 3
Tn (s)
FIGURE E5.2 Displacement, velocity, and acceleration response spectra for El Centro earthquake.
The response spectra results are shown in Figure E5.2; for damping factors of 0.02, 0.05, and
0.1, top to bottom.
EXAMPLE 5.2
Use the MATLAB script provided in Chapter 4, Example 4.13 to draw the response spectrum of the
displacement, velocity, and acceleration for an SDOF system subjected to the 1985 Mexico City
earthquake ground acceleration for 2%, 5%, and 10% damping.
Solution
Input ground acceleration is given in Figure E5.3.
0.2
Acceleration (g)
0.1
–0.1
–0.2
0 10 20 30 40 50 60 70
t (s)
FIGURE E5.3 Seismic (acceleration) time-history for 1985 Mexico City earthquake, SCT1 (EW).
Response Spectrum Analysis of SDOF System 127
Sa (g)
1
0
0 0.5 1 1.5 2 2.5 3
300
200 ζ = 0.02
Sv (in/s)
ζ = 0.05
ζ = 0.1
100
0
0 0.5 1 1.5 2 2.5 3
100
Sd (in)
50
0
0 0.5 1 1.5 2 2.5 3
Tn (s)
FIGURE E5.4 Seismic ground motion and response spectra for Mexico City earthquake.
The response spectra results are shown in Figure E5.4; for damping factors of 0.02, 0.05, and
0.1, top to bottom.
As can be seen from these spectra in Figure E5.4, the response is largest for a natural period of
∼2 s, which is the dominant period in Mexico City as discussed in Chapter 2.
Note that the relationship between the relative displacement and total acceleration mut (t ) = mωn2u(t )
is only valid for systems without damping. For this reason, the acceleration determined from this
relationship is known as the pseudo-acceleration, A, in order to distinguish A from the actual accel-
eration. That is, consider the equation of motion in terms of the relative quantities,
Since most structures have small damping (see Table 3.3), we can assume that the velocity term
is negligible (2ζωnu ωn2u ); thus, Equation 5.1 can be rewritten as
This results in the definition of the spectral pseudo-acceleration, A, in terms of the spectral
displacement, D, as
A = ωn2 D (5.4)
128 Introduction to Earthquake Engineering
0.8 ζ = 0.05
Pseudo-acceleration
0.6
Actual acceleration
Sa (g)
0.4
0.2
0
0 0.5 1 1.5 2 2.5 3
Tn (s)
FIGURE 5.5 Comparison of actual and pseudo-acceleration response spectra for El Centro earthquake.
1 2 1
kumax = kD 2 (5.5)
2 2
And if there are no other processes dissipating energy, such as damping, and so on (although not
entirely correct, we assume damping is negligible because it is small), then all the strain energy is
transferred into kinetic energy:
1 2 1
mumax = mSv2 (5.6)
2 2
From the principle of conservation of energy, these energy quantities must be equal; that is,
1 2 1
kD = mSv2 (5.7)
2 2
This results in the definition of the spectral pseudo-velocity, V, in terms of the spectral displace-
ment, D, as
k
D = V ⇒ ωn D = V (5.8)
m
Note that in general V ≠ umax, but it is sufficiently close for most practical purposes. Figure 5.6
shows the pseudo-velocity and actual velocity response spectra due to El Centro with 5% damping,
though the agreement between the response spectra is not as accurate as the comparison between
acceleration response spectra.
From Equations 5.4 and 5.8 it can be seen that
A
ωn D = V = (5.9)
ωn
Response Spectrum Analysis of SDOF System 129
40
30
Sv (in/s)
20
ζ = 0.05
10 Pseudo-velocity
Actual velocity
0
0 0.5 1 1.5 2 2.5 3
Tn (s)
or,
2π T
D =V = n A (5.10)
Tn 2π
which means that D, V, and A contain the same information; thus, we can combine the three
quantities into a single three-plot (tripartite) graph as a function of natural period, Tn, or natural
frequency, ωn.
Consider separately the relationships between D and V and that between V and A in Equation 5.9,
and now take the logarithms of both sides of each equation,
These represent equations of straight lines: the displacement with a slope of +1 and the accel-
eration with a slope of –1. That is, lines of constant D and A are inclined at +45° and –45°, respec-
tively. Thus, to construct a tripartite graph, we need to plot V versus ωn on vertical and horizontal
logarithmic scales and add logarithmic scales inclined at +45° and –45° with respect to the vertical
axis. So, V is read on the vertical scale, D is read on the scale at –45°, and A is read on the scale
at +45°. When the argument is Tn, D is read from the scale at +45° and A is read from the scale at
–45° (see Figure 5.7).
EXAMPLE 5.3
Given the structure shown in Figure E5.5, determine the response (D, V, and A) for the El Centro
ground motion using the tripartite spectra in Figure 5.7.
Tn = 0.5 s
ζ = 2%
h
100
50 V = 35 in/s
10
10
0
50
20
5
Pseudo-velocity V, in/s
10
g
1.2
10
1
5
=
5 A
Ps
n.
eud
,i
D
D
1
o- =
0.
n
ac 3.
io
ce 0
at
in
0.
2 le
m
5
ra
r
fo
tio
01 D e
n
A,
1 g
0.
0.
1
0.5
1
0.
00
01
0.
0.2
0.02 0.05 0.1 0.2 0.5 1 2 5 10 20 50
Natural vibration period Tn (s)
FIGURE 5.7 Response spectra for El Centro quake for ζ = 0%, 2%, 5%, 10%, and 20%, top to bot-
tom. (Reproduced from Chopra, A. K., Dynamics of Structures: Theory and Applications to Earthquake
Engineering, 4th edition, Prentice-Hall, Upper Saddle River, NJ, 2012. By permission of Prentice-Hall,
Upper Saddle River, NJ.)
Solution
Enter the 2%-damping response spectrum in Figure 5.7 with natural period of 0.5 s, and read
V = 35 in/s from the vertical scale, D = 7.5 in from the scale at +45°, and A = 1.2g from the scale
at –45°. The lines showing these values are depicted in Figure 5.7.
EXAMPLE 5.4
The reinforced concrete bridge structure shown in Figure E5.6 is subjected to the El Centro earth-
quake; determine the total stiffness, structural period, deflection of the deck, and base shear. The
deck weighs 750 kips and the substructure has 2% damping, modulus of elasticity, E = 3000 ksi,
and the other geometric properties shown in Figure E5.6.
Fixed–pinned Fixed–pinned 15 ft
20 ft
I = 25,000 in4 I = 20,000 in4
Fixed–fixed
30 ft
I = 60,000 in4
Solution
1. Mass, stiffness, and natural period of the system: The bridge can be modeled as an SDOF
system assuming a rigid deck and only lateral deformations of the columns occur. The
stiffness of the system is then the sum of column lateral stiffnesses. The mass and stiffness
of the SDOF system are calculated as follows:
Mass,
w 750 kips
m= = = 1.94 kips s2 /in
g 386.4 in /s2
Stiffness (see Table 3.1), the 30-foot column is fixed–fixed, while the other two are
fixed–pinned,
m 1.94 kips s2
Tn = 2π = 2π = 0.906 s
k 93.4 kips /in
D ≅ 6 in
3. Determine the maximum base shear: Calculate the maximum base shear at the support
using the force–displacement relationship or equilibrium as described in Figure 3.24:
A careful examination of the response spectra shown in Figure 5.7 shows that there are regions
where each of the response parameters (D, V, and A) are nearly constant over a wide range of values
for Tn. For example, for very flexible systems (those with large periods), the displacement is equal
to the peak ground displacement for all damping ratios and experience negligible accelerations.
However, stiff systems (those with small periods) accelerate at the PGA for all damping ratios and
experience negligible relative displacements (only ground displacement). Also, damping has the
greatest effect over the center region. This analysis of Figure 5.7 is specific to a set of response spec-
tra for a specific earthquake. In order to compare the results of different earthquakes, we normalize
each response parameter (spectral displacement, velocity, and acceleration) with respect to its cor-
responding peak ground value (peak ground spectral d isplacement, peak ground velocity, and PGA,
respectively) as shown in Figure 5.8.
Figure 5.8 also depicts an idealized response spectrum, shown by a dashed line. We can observe
that for the first portion (Tn < 0.5 s) of the response spectra, the acceleration is largest (acceleration
sensitive), before decreasing precipitously; for the middle portion (0.5 s < Tn < 3 s), the velocity is
largest (velocity sensitive); and for the last portion (Tn > 3 s), the displacement is largest (displace-
ment sensitive). This phenomenon is one of the reasons for determining V and A. Furthermore, as
132 Introduction to Earthquake Engineering
10
Acceleration Velocity Displacement
0 sensitive sensitive sensitive
5 10
10
2 c d
10
1
Td = 3.0 s
Tc = 0.5 s
e
1
b
0.5
V/ugo
1
.
Ta = 0.035 s
0.
A/
1
..
0.2
ug
go
1
/u
Tb = 0.125 s
0.
Te = 10 s
0.1
0.
Tf = 15 s
a 01
01
0.05
0.
0.
00
1
0.02
0.02 0.05 0.1 0.2 0.5 1 2 5 10 20 50
Tn (s)
FIGURE 5.8 Spectral regions for El Centro quake; ζ = 5%. (Reproduced from Chopra, A. K., Dynamics of
Structures: Theory and Applications to Earthquake Engineering, 4th edition, Prentice-Hall, Upper Saddle
River, NJ, 2012. By permission of Prentice-Hall, Upper Saddle River, NJ.)
mentioned previously, delineating regions along the spectra where the different response quantities
are dominant (or maximum) is useful in establishing design spectra, as discussed in Section 5.1.1.
10
5 0
10
Mean + 1σ
10
2
go
/u
c d
D
10
A/
..
1 ug o Mean
Tf = 33 s
Tc = 0.349 s
1
b
0.5
V/ugo
1
.
e
Ta = 1/33 s
Td = 3.135 s
0.
0.2
1
1
0.
Tb = 1/8 s
Te = 10 s
0.1 f
0.
01
a
0.05
01
0.
0.
00
0.02
1
FIGURE 5.9 Mean and mean + 1σ spectra for a group of 10 quakes; ζ = 5%. (Reproduced from Chopra,
A. K., Dynamics of Structures: Theory and Applications to Earthquake Engineering, 4th edition, Prentice-
Hall, Upper Saddle River, NJ, 2012. By permission of Prentice-Hall, Upper Saddle River, NJ.)
3. Determine the mean and standard deviation (σ) at each period. Figure 5.9 schematically
shows the probability distributions for V at Tn = 0.25, 1, and 4 s, which also show that the
coefficient of variation (COV = σ/mean) changes with Tn.
Figure 5.9 also clearly shows that the graphs of the mean and the mean + 1σ are smoother than
any individual parent spectrum, one of which is Figure 5.7.
The elastic design spectrum shown in Figure 5.10 is then obtained by smoothing out into a
series of straight lines the spectra graphs. The dashed lines in Figures 5.9 and 5.10 represent the
means of the response spectra. This simplified empirical approach to constructing elastic response
spectra was first developed by Professors Nathan Newmark and William Hall, and is known as
the Newmark–Hall method. The key parameters are the structural amplification factors, α, which
represent the spectral amplification above ground motion. A summary of structural amplification
factors is listed in Table 5.1. These empirical values were computed using records from firm ground,
such as rock, soft rock, and competent sediments. Though limited in its applicability because of the
small number of records and narrow site characteristics, many pieces of the general shape can be
found in current code spectra, as will be demonstrated in Chapter 8.
After selecting a damping ratio, ζ, and a mean or mean + 1σ plot, the general procedure for
drawing a design response spectrum entails the following steps:
1. Mark the key periods delineating the transition regions, Ta, Tb, Te, and Tf. The periods
delineating the velocity, acceleration, and displacement-controlled regions are obtained
from the intersection of A and V and the intersection of V and D.
2. Plot lines corresponding to values of ugo, u go , and ügo for the specified design ground
motion. This peak ground motion graph represents the backbone of the response spectrum
and can be considered the lower bound of the response.
134 Introduction to Earthquake Engineering
Tc αV ugo Td
α
Du
go
Te
go
Au
ugo
α
Tb u
go
V (log scale)
go
u
Tf
Ta
1/33 s 1/8 s 10 s 33 s
Tn (log scale)
TABLE 5.1
Structural Amplification Factors for Elastic Design Spectra
Mean (50th Percentile) σ (84.1th Percentile)
Damping (%) αA αV αD αA αV αD
2 2.74 2.03 1.63 3.66 2.92 2.42
5 2.12 1.65 1.39 2.71 2.30 2.01
10 1.64 1.37 1.20 1.99 1.84 1.69
ζ 3.21–0.68 lnζ 2.31–0.41 lnζ 1.82–0.27 lnζ 4.38–1.04 lnζ 3.38–0.67 lnζ 2.73–0.45 lnζ
Source: Newmark, N. M. and W.J. Hall, Earthquake Spectra and Design, Earthquake Engineering Research Institute,
Oakland, CA, 1982.
3. Obtain values of α A, αV, and αD for a specified damping ratio, ζ, from Table 5.1; where the
mean or 50th percentile represents a 50% chance of being exceeded (50% exceedance).
4. Amplify the ground peak values (backbone) using these parameters to get: α Aügo, αV u go ,
and αDugo.
5. Finally, connect the transition lines from a to b and from e to f.
The axes of the Newmark–Hall elastic design spectrum were normalized with respect to ügo = 1g,
u go = 48 in/s, and ugo = 36 in. Values of ügo can be obtained from deterministic or probabilistic site
hazard analysis, whereas ugo and u go can be determined from site hazard analyses or as empirical
functions of ügo when only ügo is available. The following empirical relationships are recommended
for estimating u go and ugo in this case:
A number of other researchers have conducted statistical analyses and determined values
that vary considerably because these values depend on magnitude, epicentral distance, and soil
characteristics at the recording station site.
EXAMPLE 5.5
Draw the 84.1th percentile response spectrum for 5% damping using the tripartite graph shown
in Figure 5.11. The ground motion is anticipated to have a PGA of ügo = 0.5g.
Solution
1. Determine the peak ground velocity, u go , and peak ground displacement, ugo: Use the
PGA and Equation 5.13 to determine these quantities.
Peak ground velocity,
in /s in /s
u go = 48 ugo = 48
⋅ 0.5 g = 24 in /s
g g
ugo = 6u go
2
/ugo = 6( 24 in /s)2 / 0.5(386 in /s2 ) = 18 in
Plot the backbone quantities on the tripartite graph paper provided in Figure 5.11, as
shown in Figure E5.7.
2. Determine the structural amplification factors, αA, αV, and αD: For 84.1th percentile spec-
trum and 5% damping, these quantities are obtained from Table 5.1 as
100
50
0
.0
10
20
10
0.
00
10
00
1.
10
V (in/s)
5
,i
.0
D
0
10
A,
0.
g
2
1.
00
1
01
0.
0.
0.5
10
00
0.
0.2
0.
01
100
55.2 in/s
50
36
.2
0
in
24 in/s
.0
10
g
36
20
1.
10
18
0.
in
00
10
5g
00
0.
1.
10
V (in/s)
n
5
.0
,i
0
D
10
A,
0.
2 g
1.
00
1
01
0.
0.
10
0.5
00
0.
0.2
0.
01
FIGURE E5.7 Elastic design response spectrum for peak ground acceleration of 0.5g and 5% damping.
A = ügoαA = 0.5g (2.71) = 1.36g
V = u go αV = 24 in/s (2.3) = 55.2 in/s
D = ugoαD = 18 in (2.01) = 36.2 in
Plot these quantities in the same graph as the backbone quantities, Figure E5.7.
4. Complete the response spectrum by connecting the transition lines. Alternatively,
the spectral graph can be obtained using a conventional graphing computer program
such as MATLAB.
Note that the results of this example would be the same if we had derived the response spectrum
for the standard peak ground values of ügo = 1g, u go = 48 in/s, and ugo = 36 in and scaled the result-
ing spectrum by a factor of 0.5. Therefore, constructing an elastic design response spectrum using
the Newmark–Hall method only requires the damping ratio and expected PGA, ügo = ηg; The scal-
ing factor, η, can easily be applied to the ordinates of the design spectra.
We have used the tripartite graph to delineate the three main regions of elastic design spectra
(acceleration, velocity, and displacement sensitive); however, for design purposes, it is more conve-
nient to return to the single acceleration spectra. This can be accomplished by plotting Equation 5.10
as shown in Figure 5.12 for the Newmark–Hall spectrum, ügo = 1g, u go = 48 in/s, and ugo = 36 in
and ζ = 5%. The axes of this spectrum are also plotted using linear scales in Figure 5.13. This figure
is similar in shape to contemporary code-based design spectra.
In cases where only a single value of the pseudo-acceleration, A, is needed, it may not be necessary
to construct the entire elastic design response spectrum; particularly if the period is between 1/8
and 10 s (constant acceleration, velocity, and displacement regions). For such a case, the governing
design acceleration is given by the smallest of the following quantities:
Response Spectrum Analysis of SDOF System 137
2.71
11.7Tn0.704
1 1.8Tn–1
A (g)
7.4Tn–2
0.1
0.01
0.01 1/33 0.1 1/8 0.66 1 4.12 10
Tn (s)
FIGURE 5.12 84.1th percentile acceleration response spectrum for ζ = 5%, plotted using logarithmic scales.
A = ügoα A
A = ωnV = 2πu go αV/Tn
A = ωn2D = (2π/Tn)2ugoαD
Also, as noted earlier, spectral quantities not only depend on PGA magnitude, but also on soil
characteristics at the recording station site and epicentral distance. The values for α A, αV, and αD in
the Newmark–Hall method were derived based on a limited number of records that were obtained
on a firm ground. More comprehensive studies have revealed that local site soil conditions have
3
2.71
2.5
2
11.7Tn0.704
A (g)
1.5 1.8Tn–1
0.5
0
0 1/8 0.5 0.66 1 1.5 2 2.5 3
1/33
Tn (s)
FIGURE 5.13 84.1th percentile acceleration response spectrum for ζ = 5%, plotted using linear scales.
138 Introduction to Earthquake Engineering
3
A B Soft to medium clay and
sand—15 records
C Stiff soils
(<200 ft)—31 records
1
D
Rock – 28 records
0
0 0.5 1.0 1.5 2.0 2.5 3.0
Tn
FIGURE 5.14 Influence of local soil conditions on response spectra for 5% damping. (Based in part on Seed,
H. B. and I. M. Idriss, Ground Motions and Soil Liquefactions during Earthquakes, Earthquake Engineering
Research Institute, Oakland, CA, 1982.)
a significant impact on spectra shapes. As shown in Figure 5.14, for soft soils, ügo remains the
same or decreases relative to firm soil, but u go and ugo increase. Also, as the epicentral distance
increases, the geology over the distance tends to filter the acceleration effects of certain frequencies.
This results in the short-period portion of the design spectra being controlled by nearby earthquakes
and the long-period portion of the design spectra controlled by earthquakes at further distances
from the structure.
Contemporary design spectral analysis is based on a uniform hazard spectrum that accounts
for the effects of all earthquakes expected at a site over a period of time using probability-based
relations known as attenuation relations. These account for PGA magnitude, local site soil con-
ditions, and epicentral distance all together; and along with probabilistic models of earthquake
occurrence, these relations are used to obtain site-specific design spectra. Usually, however, the
entire spectra are not computed following this procedure; rather, only accelerations correspond-
ing to a couple of natural periods (0.2 and 1 s) are determined, which are then used to gener-
ate full design spectra using approximate relations. This approach will be discussed further in
Chapter 8.
EXAMPLE 5.6
For the concrete bridge structure of Example 5.4, determine the deflection of the deck and base
shear when the bridge is subjected to ground acceleration due to an earthquake characterized by
the 84.1% design spectrum scaled to 0.32g PGA. The deck weighs 750 kips and the substructure
has 5% damping, modulus of elasticity, E = 3000 ksi, and the other geometric properties are
given in Example 5.4.
Solution
1. Mass, stiffness, and natural period of the system: The bridge can be modeled as an SDOF
system assuming a rigid deck and considering only lateral deformations of the columns.
The stiffness of the system is then the sum of column lateral stiffnesses. The mass and
stiffness of the SDOF system are calculated in Example 5.4; the resulting quantities are
V 35.3 in /s
D= = = 5.1in
ωn 18.38 rad /s
3. Determine the maximum base shear: Calculate maximum base shear at the support
using the force–displacement relationship or e quilibrium as described in Figure 3.24:
EXAMPLE 5.7
The building frame shown in Figure E5.8, is subjected to a ground acceleration from an earth-
quake characterized by the 84.1% design spectrum scaled to 0.25g PGA, determine (i) maxi-
mum displacement u, (ii) the maximum base shear, and (iii) the maximum normal stresses in the
columns (Ix = 75 in4 and Sx = 18 in3 for W8x21). Let us assume that the beam is rigid and system
damping is 5%.
Solution
1. Determine mass, stiffness, and natural period and frequency of the SDOF system: Since we
have assumed the beam to be rigid, the stiffness of the system is given by the sum of column
lateral stiffnesses. The mass and stiffness of the SDOF system are calculated as follows:
Mass,
W 15 kips(1000 lb /kip)
m= = = 38.86 lb s2 /in
g 386.4 in /s2
15 kips
15 ft
m 38.86 lb s2 /in
Tn = 2π = 2π = 0.828 s
k 2237.6 lb /in
k 2237.6 lb /in
ωn = = = 7.59 rad /s
m 38.86 lb s2 /in
2. Determine the maximum displacement, D: To determine D due to an 84.1% design
spectrum, scaled to 0.25g PGA, first, enter the response spectrum in Figure 5.12 with
Tn = 0.83 s > 0.66 s, so velocity is constant,
V 27.6 in/s
D= = = 3.6 in
ωn 7.59 rad /s
3. Determine the maximum base shear: Calculate maximum base shear at the support for
each column using the force–displacement relationship:
3EI
3( 29, 000 ksi)(75 in4 )
Vmax = kuo = 3 uo = 3.6 in = 4.1kips
L (15 ft ×12 in /ft )3
4. Maximum normal stress due to bending: The maximum bending moment at the top of
each column is calculated from equilibrium of the column member. Since the column has
a pin support, summing moments about the top of the column will result in the following
maximum internal bending moment (see Figure 3.24):
Fs
Corresponding linear system
Fo Fo = kuo
Fy Elastoplastic system
Fy = kuy
u
uy uo um
–Fy
We can also define a parameter based on the fictitious corresponding linear system and the yield
force, known as the yield strength reduction factor, Ry:
Fo kuo uo
Ry = = = (5.16)
Fy kuy uy
where uo represents the maximum displacement of the corresponding linear elastic system.
Note that when u ≤ uy, Equation 5.16 is related to a real elastic system, whereas for u > uy, the
system behaves inelastically and may not return to its equilibrium position after the shaking has
stopped—it may experience permanent deformation.
In addition to μ, we can define a second displacement ratio between maximum displacement, um,
and the maximum displacement of the fictitious corresponding linear system, uo, which relates Ry and μ,
um um uy μuy μ
= = = (5.17)
uo uo uy uo Ry
This gives the ductility demand imposed on an elastoplastic system by ground motion:
um
μ= Ry (5.18)
uo
142 Introduction to Earthquake Engineering
50
10
10
ζ = 2, 5, 10%
0
20
6g
0.1
=
10 1 Ay
10
5
Vy (in/s)
1 µ=1
0.
1
2
01
1
0.
0.
1
µ=4
Ay
0.5
,g
n
y,i
1
00
0.
µ=8
u
01
0.
0.2
0.02 0.05 0.1 0.2 0.5 1 2 5 10 20 50
Tn (s)
FIGURE 5.16 Elastoplastic response spectra for El Centro quake. (Reproduced from Chopra, A. K.,
Dynamics of Structures: Theory and Applications to Earthquake Engineering, 4th edition, Prentice-Hall,
Upper Saddle River, NJ, 2012. By permission of Prentice-Hall, Upper Saddle River, NJ.)
For the purpose of designing a system, ductility capacity must be larger than the ductility
demand, that is, the system must have the ability to deform beyond the elastic limit.
The comparison between the elastoplastic and the corresponding linear systems is also useful in
establishing a procedure for using the elastic design response spectrum in Figure 5.10 to c reate the
inelastic design response spectrum. Data presented by Chopra (2012) shows that in the displacement
and velocity-sensitive regions, the maximum displacement is nearly equal to the ground displacement
and the strength reduction factor is nearly equal to the ductility factor, that is, the result is independent
of inelastic response. However, in the acceleration-sensitive region, the ductility demand is much
greater than the strength reduction factor; short-period systems are stiffer and tend to attract more load,
requiring a larger stiffness. Thus, we now have the response quantities (yield deformation, pseudo-
velocity, and pseudo-acceleration) being functions of three different variables (period, damping ratio,
and ductility ratio) (see Figure 5.16). This figure clearly indicates that d amping has a decreasing effect
as ductility increases and is much less effective than d uctility in reducing the maximum response.
Furthermore, to increase a system’s earthquake resistance, it can be designed by either making it
strong or ductile (or a combination of the two), because i ncreasing ductility greatly decreases response.
Given yield displacement, uy, the relationships between elastic (Equations 5.9) and inelastic
quantities are as follows:
Dy = uy
Vy = ωnDy = ωnuy
Ay = ωn2Dy = ωn2uy = (2π/Tn)2Dy
And since um = μuy or um = μDy,
T 2
um = µ n Ay (5.19)
2π
EXAMPLE 5.8
Consider the building frame and characteristics given in Example 5.7. If the building frame is
subjected to the El Centro earthquake, determine the maximum required column strength for a
ductility ratio of 4.
Response Spectrum Analysis of SDOF System 143
Solution
1. Mass, stiffness, and natural period and frequency of the SDOF system were deter-
mined in Example 5.7. That is, m = 38.86 lb s2/in, k = 2237.6 lb/in, Tn = 0.828 s, and
ωn = 7.59 rad/s.
2. Determine the required column strength for a ductility demand of 4, Fy: To determine Fy,
first, enter the response spectrum in Figure 5.16 with Tn = 0.83 s and read the pseudo-
acceleration, Ay,
Ay ≅ 0.16g
um kFe F
μ= = = e (5.20)
uy kFy Fy
Now we can rewrite this relationship in terms of accelerations using the system mass,
Fe Fe muy uy
μ= = = = (5.21)
Fy Fi mui ui
Fs
Fs
Fe Fe
Fy Fi
Fy Fi
u u
uy um uy ue ui
FIGURE 5.17 Force deformation curves for elastic and elastoplastic systems.
144 Introduction to Earthquake Engineering
Since Fi = Fy. This results in the maximum inelastic acceleration being equal to the maximum
elastic acceleration divided by the ductility factor,
uy A
ui = ⇒ Ay = (5.22)
μ μ
The result presented in Equation 5.22 agrees relatively well with experimental data for the
c onstant displacement and velocity regions, but not for the constant acceleration region. For the
constant acceleration region, the relationship is obtained using the equivalence of energy in
the elastic (AREAe) and inelastic (AREAi) systems. AREAe is the hatched area, while AREAi is
the solid area in the right-hand-side graph of Figure 5.17. That is,
1 1 u2
AREAe = ue Fe = Fy e (5.23)
2 2 uy
1
AREAi = uy Fy + (ui − uy )Fy (5.24)
2
1 ue2 1
Fy = uy Fy + (ui − uy )Fy (5.25)
2 uy 2
Using the definition of the ductility ratio, μ = ui/uy, we can now solve for the inelastic
displacement in terms of the elastic displacement,
μ μ
ui = ue ⇒ Dy = D (5.26)
2μ − 1 2μ − 1
ue A
ui = ⇒ Ay = (5.27)
2μ − 1 2μ − 1
Tc V = αV ugo Td
D
=
α
Du
Tc′ Vi = V/µ Td′ go
u go
Te
Di
αA
) 0.5
=
=
A
D/
µ–
µ
(2
A/
Tb
=
A
i
V or Vi (log scale)
Tf
D/
µ
u go
D/
Ta µ
u go
1/33 s 1/8 s 10 s 33 s
Tn (log scale)
Thus, to obtain the maximum displacement, we use the relationship between displacement and
pseudo-acceleration:
T 2
uo = µ n Ay (5.28)
2π
Also, the maximum force (required yield strength) can be determined using the relations between
mass and acceleration:
Fo = mAy (5.29)
EXAMPLE 5.9
Consider the building frame shown in Figure E5.9 with ζ = 5% and Ix = 75 in4 for the two columns
subjected to a ground acceleration due to an earthquake characterized by the 84.1% design spec-
trum scaled to 0.5g PGA. Assume elastoplastic behavior and determine the lateral deformation
and force for which the frame should be designed if μ = 4.
1500 lbs
15 ft
Solution
1. Mass, stiffness, and natural period of the SDOF system: Since we assumed the beam to
be rigid, the stiffness of the system is given by the sum of column lateral stiffnesses. The
mass and stiffness of the SDOF system are calculated as follows:
Mass,
W 1500 lbs
m= = = 3.88 lb ⋅ s2 /in
g 386.4 in /s2
m 3.88 lb s2 /in
Tn = 2π = 2π = 0.262 s
k 2237.6 lb /in
2. Determine the maximum displacement for ductility of 4, uo: First determine A due to an
84.1% design spectrum, scaled to 0.5g PGA, by entering the response spectrum in Figure
5.12 with Tn = 0.26 s, which is in the acceleration-sensitive region,
The maximum displacement is computed using Equation 5.28 and Figure 5.18,
2
T 2 T 2 A 0.262 s 1.335(386.4 in /s2 )
uo = µ n Ay = µ n = 4 = 1.353 in
2π 2π 2µ − 1 2π 2(4) − 1
3. Determine the required column strength for a ductility of 4, Fy: To determine Fy, use
A = 1.335g and Equation 5.29.
PROBLEMS
5.1 Use the MATLAB script provided in Chapter 4, Example 4.13 to draw response
spectra for the displacement, velocity, and acceleration for an SDOF system sub-
jected to the 1994 Northridge earthquake ground acceleration for 2%, 5%, and 10%
damping.
5.2 The building frame shown is subjected to the El Centro earthquake; determine the
total stiffness, structural period, deflection of the beam, base shear, and bending
stresses in each of the columns. The roof weighs 10 kips and the substructure has 5%
damping, modulus of elasticity, E = 29,000 ksi, and the other geometric properties
shown.
Response Spectrum Analysis of SDOF System 147
10 kips
5.3 The following structural system is subjected to the El Centro earthquake; determine
the total stiffness, structural period, deflection of the beam, and base shear. The sub-
structure has 2% damping, modulus of elasticity, E = 3000 ksi, and the other geometric
properties as shown.
W = 750 kips u
5.4 A 25 kip sensitive piece of equipment will be installed on the roof of the following
frame shown with a rigid slab and four columns with EIx = 125,000 kip in2 and 5%
damping. The frame is located in Mexico City at a site with response spectrum depicted
in Example 5.2. Determine approximately the magnitude of the shear force (Vreq’d)
needed to design the equipment anchorage.
25 kips
Vreq’d
12 ft
5.5 Use a tripartite graph to draw the elastic 50th percentile design response spectrum with
a maximum ground acceleration of 0.25g and a damping ratio of 2%.
5.6 For the frame of Problem 5.2, determine the deflection of the beam, base shear, and
bending stresses when the frame is subjected to ground acceleration due to an earthquake
characterized by the 84.1% design spectrum scaled to 0.5g peak ground acceleration. The
substructure has 5% damping and the other geometric properties given in Problem 5.2.
5.7 Burn’s tower (a water tank weighing 4200 kips) on the campus of University of the
Pacific is supported on a 120-ft high cantilever tower with a stiffness of 2000 kips/in
and 5% damping. Determine the design values for lateral deformation and base shear if
the tower is subjected to ground acceleration due to an earthquake characterized by the
84.1% design spectrum scaled to 0.3g peak ground acceleration.
148 Introduction to Earthquake Engineering
5.8 Use a tripartite graph to draw the inelastic 50th percentile design response spectrum
with a maximum ground acceleration of 0.25g, a damping ratio of 2%, and a ductility
ratio of 5.
5.9 Given the following 14-ft. steel tower subjected to a ground acceleration due to an earth-
quake characterized by the 84.1% design spectrum scaled to 0.25g peak ground accel-
eration, determine (i) maximum lateral displacement, and (ii) the maximum base shear.
Assume system damping of 5% and ductility of 5.
W = 38.6 kips
Ix = 75 in4
Sx = 18 in3
5.10 Consider the following building frame with a rigid beam that is pinned con-
nected to one column and rigidly connected to the other as shown; each column has
EI = 40,000,000 kip in2. The frame is subjected to ground acceleration due to an
earthquake characterized by the 84.1% design spectrum scaled to 0.33g peak ground
acceleration. Determine maximum stress in each column given that Sx = 2,000 in3 and
damping is 5%.
200 kips
10 ft Pinned
15 ft
5.11 Solve for the maximum base shear in the frame shown in problem 5.10 assuming a duc-
tility ratio of 4.
5.12 Consider a one-story frame with lumped weight w and natural vibration period,
T = 0.25 s, in the elastic range. Determine the lateral deformation and force in terms of
w for which the frame should be designed if μ = 6. Assume that ζ = 5%, elastoplastic
behavior, and an 84.1% design earthquake with ügo = 0.25g.
uo
Fo
w
k/2 k/2
5.13 Two identical power transformers to be erected adjacent to each other are supported
by steel (E = 29,000 ksi) pedestals and can be idealized as SDOF systems. The ped-
estals are designed to withstand inelastic deformations of up to six times their yield
Response Spectrum Analysis of SDOF System 149
values. Using an inelastic 50th percentile design response spectrum with a maximum
ground acceleration of 0.25g, estimate the minimum clear distance (Δ) that must be left
between the two transformers to avoid collision. I = 1380 in4 and W = 1.5 kips.
W W
∆
20 ft
REFERENCES
Chopra, A. K., Dynamics of Structures: Theory and Applications to Earthquake Engineering, 4th edition,
Prentice-Hall, Upper Saddle River, NJ, 2012.
Seed, H. B. and I. M. Idriss, Ground Motions and Soil Liquefactions during Earthquakes, Earthquake
Engineering Research Institute, Oakland, CA, 1982.
6 Generalized SDOF
System Analysis
After reading this chapter, you will be able to:
Before presenting the formulation for a system discretized into multiple degrees of freedom, it
is instructive to study the development of an approach that condenses these MDOF systems into a
SDOF case. Unlike the MDOF system formulation, this approach does not require simultaneous
manipulation of multiple algebraic equations. The key to obtaining relatively accurate results with
a generalized SDOF system analysis is selecting an appropriate shape function for the deformed
system during vibration.
In this chapter, we also describe the analysis of systems with distributed mass and stiffness as
generalized SDOF systems. While this analysis approach is not as precise as modeling a system
with an infinite number of degrees of freedom, modeling systems as generalized SDOF systems
is more accurate than approximating the entire system as a simple SDOF system as discussed in
Chapter 3. Furthermore, utilizing an infinite number of degrees of freedom to model the distributed
mass and stiffness as a function of position would require formulation in terms of partial differential
equations in both position and time.
The generalized SDOF system analysis presented in this chapter is based on replacing the
required partial differential equation in time and space with an approximate differential equation in
time only. The space variation is tracked using an approximate equation of the mode of vibration for
the system. The process entails selecting a point of interest along the distributed system and track-
ing its motion with respect to time using an SDOF analysis; in order to obtain the response of all
other points, we scale the magnitude of the point in question based on the approximate equation of
the mode of vibration. The solution to the differential equation representing the generalized SDOF
system follows the standard procedures presented in Chapter 3; and we can apply all the solution
techniques discussed thus far, including the response spectrum analysis presented in Chapter 5.
Replacing the distributed dynamic properties (mass, damping, and stiffness) with lumped prop-
erties at the point being tracked requires integrating the variation in these properties along the
length of the system. For discrete MDOF systems, the process involves combining the discrete
properties associated with each DOF into a set of generalized properties. These generalized proper-
ties are most useful in developing the participation factor, which is critical in applying the results
from an SDOF system analysis to an MDOF system analysis.
The formulation of the equation of motion of a generalized SDOF system entails condensing
all degrees of freedom into one. In the following sections, we present analysis procedures for a
multistory shear building frame (discrete system) and a cantilever tower with distributed properties
(continuous system), both subjected to earthquake ground motions. In each case, we assume axial
deformations to be negligible during the dynamic excitation, which leaves the lateral translation
(or deflection) as the only possible deformation of the system.
151
152 Introduction to Earthquake Engineering
6.1.1 Equation of Motion
The actual motion of the shear building shown in Figure 6.1 can be obtained by determining the
motion of the top of the building, z(t), and scaling it to the other levels using an appropriate shape
vector, ψj. That is, at an arbitrary level, j, the relative displacement is
u j (t ) = ψ j ⋅ z (t ) (6.1)
where:
j = 1, 2,… , N
The motion at the top of the building, z(t), can be obtained using a generalized equation of motion
for the equivalent SDOF system shown in Figure 6.1. This equation of motion can be obtained
using D’Alembert’s principle as discussed in Chapter 3 or the principle of virtual work, which
is the approach we follow in this chapter. The resulting equation is of the same form as the one
ut
mN z
N uN δz
z(t)
kN
~
m
mj
j uj δuj
kj
~
k
m2
u2
2
k2
m1
1 u1
k1
ug(t)
ug(t)
FIGURE 6.1 MDOF shear building and equivalent generalized SDOF system.
Generalized SDOF System Analysis 153
that c haracterizes the vibration of the SDOF oscillator introduced in Chapter 3; thus, all analysis
procedures covered so far are applicable to this case.
The shape vector, ψj, can be obtained from any function, ψ(x), provided it satisfies the geometric
boundary conditions. However, the accuracy of the results depends on how close the assumed shape
function approximates the actual shape of the deformed frame. Also, the function ψ(x) must be
normalized at the top of the building such that ψ(H) = 1.0, where H is the height of the building.
Figure 6.2 presents several possible shape functions based on building elevation aspect ratio.
To account for the ground excitation loading in the formulation of the equation of motion, we
also need the total displacement at level j, which is a combination of the relative displacement, uj(t),
and the ground level lateral displacement, ug(t):
utj (t ) = u j (t ) + ug (t ) (6.2)
We can now use the principle of virtual work to derive the generalized equation of motion.
This principle states that the virtual work of all internal forces, δWI, equals the virtual work of all
external forces, δWE, for a given virtual displacement, δu. That is,
where δWI includes the contribution from stiffness and damping forces:
And δWE includes the contribution from inertial forces of each mass:
The virtual work of each of these groups of forces is given by summation of the product of the
magnitude of the force and the corresponding virtual displacement. The stiffness and damping
forces can be viewed as effective shear forces in each story, whereas the inertial forces can be
regarded as effective lateral loads applied at each level. The stiffness forces are functions of story
D
1
D
1
H
D
1
H
ψ(x)
H
ψ(x) ψ(x)
Low-rise, H/D < 1.5 Mid-rise, 1.5 < H/D < 3 High-rise, H/D > 3
πx πx
ψ(x) = sin ( 2H ) ψ(x) = ( Hx ) ψ(x) =1–cos ( 2H )
FIGURE 6.2 Possible shape functions based on building elevation aspect ratio. (With kind permission from
Springer Science+Business Media: Naeim, F. 2001. The Seismic Design Handbook, 2nd edition, Dordrecht,
The Netherlands.)
154 Introduction to Earthquake Engineering
drift (relative displacement from one floor to the next); the damping forces are functions of relative
velocity from one floor to the next; and the inertial forces are functions of the total acceleration at
each level.
First, determine the contribution of the stiffness forces to the virtual work equation. The stiffness
force at level j, fsj, is equal to the corresponding story stiffness times story drift,
fsj = k j (u j − u j −1 ) (6.6)
where kj is the sum of all the lateral stiffness systems (columns, shear walls, bracing, etc.) in the
direction in question at story j. Substituting Equation 6.1 evaluated at stories j and j–1 into this
equation yields
We can factor out z(t) and replace the shape vector values with the change in values between two
stories, (Δψj = ψj − ψj−1),
where δuj represents the internal virtual displacement of the structure. Substituting the value of fsj
from Equation 6.8 and virtual values of Equation 6.1 into Equation 6.9 results in the internal virtual
work due to stiffness,
We can then add the contribution of all N stories to obtain the total internal work from the stiff-
ness forces in each story,
δWstiffness = δ z ⋅ z(t ) ∑ k ⋅ Δψ
j =1
j
2
j (6.12)
Similarly, we can determine the contribution of the damping forces to the virtual work equation.
The damping force at level j, fdj, is equal to the corresponding story damping times story relative
velocity,
where cj is the sum of all the damping systems in the direction in question at story j.
Generalized SDOF System Analysis 155
Substituting Equation 6.1 evaluated at stories j and j–1 into this equation yields
Substituting the value of fdj from Equation 6.15 and virtual values of Equation 6.1 into Equation
6.16 results in the internal virtual work due to damping,
We can then add the contribution of all N stories to obtain the total internal work from the damp-
ing forces in each story,
δWdamping = δ z ⋅ z(t ) ∑ c ⋅ Δψ
j =1
j
2
j (6.19)
Substituting Equations 6.12 and 6.19 into Equation 6.4, we obtain the total internal virtual work,
N N
δWI = δ z ⋅ z(t ) ∑ j =1
k j ⋅ Δψ 2j + δ z ⋅ z(t ) ∑ c ⋅ Δψ
j =1
j
2
j (6.20)
Finally, we determine the contribution of the external forces to the virtual work equation. The
inertial force at level j, f Ij, is equal to the corresponding mass times story total acceleration,
Substituting the second derivative with respect to time of Equation 6.2 into this equation yields
Now taking the second time derivative of Equation 6.1 evaluated at story j and substituting into
Equation 6.22 gives,
fIj = m j (ug +
z (t )ψ j ) (6.23)
156 Introduction to Earthquake Engineering
This is negative because the sense of the force is opposite to the direction of the virtual displace-
ment. Substituting the value of f Ij from Equation 6.23 and virtual values of Equation 6.1 yields
δWinertial j = −m j (ug +
z (t )ψ j )δzψ j (6.25)
We can now sum the contribution of all N stories to obtain the total external work from the
inertial forces at each level j,
N N
We can now apply the principle of virtual work (Equation 6.3) by setting the internal virtual
work, Equation 6.20, equal to the external virtual work, Equation 6.26,
N N N N
δ z ⋅ z (t ) ∑ j =1
k j ⋅ ∆ψ 2j + δ z ⋅ z(t ) ∑j =1
c j ⋅ ∆ψ 2j = −δ z ⋅ ug ∑ j =1
m j ψ j − δ z ⋅ (
z t) ∑m ψ
j =1
j
2
j (6.27)
After eliminating the virtual displacement, δz, since it appears in each term, and rearranging this
equation we obtain
N N N N
z (t )∑ j =1
∑
k j ⋅ ∆ψ 2j + z(t )
j =1
c j ⋅ ∆ψ 2j +
z (t ) ∑ j =1
m j ψ 2j = −ug ∑m ψ
j =1
j j (6.28)
This is the equation of motion of a generalized SDOF system, which can be written in the same
form as the equation of motion derived in Chapter 3,
+ cz
mz = − Lu
+ kz g (t ) (6.29)
where:
∑
N
m = m j ⋅ ψ 2j is the generalized mass
j =1
k = ∑
N
k j ⋅ (∆ψ j )2 is the generalized stiffness
i =1
c = ∑
N
c j ⋅ (∆ψ j )2 is the generalized damping
i =1
L = ∑
N
m j ⋅ ψ j is the generalized force
i =1
N in the summation is the number of stories in the building
Δψj is the relative value of the shape function of two consecutive levels, that is, ∆ψ j = ψ j − ψ j−1
mj is the mass at level j
kj is the stiffness for the jth story
cj is the damping for the jth story
Generalized SDOF System Analysis 157
where:
the generalized natural frequency is given as
k (6.31)
ωn =
m
L
Γ = (6.32)
m
m
Tn = 2π (6.33)
k
z = Γ ⋅ D (6.34)
This maximum displacement can be distributed to the other floors using the shape vector, ψj
(Equation 6.1), as shown in Figure 6.3,
u jo = ψ j ⋅ z = ψ j ⋅ Γ ⋅ D (6.35)
Γ
u jo = ψ j A (6.36)
ωn2
158 Introduction to Earthquake Engineering
mN z mN
fNo uNo fNo
kN kN
mj mj
fjo ujo fjo
kj kj
hj fio
Vmax i
m1 Mmax i
f1o u1o hi
k1
Mmax b Vmax b
FIGURE 6.3 Maximum dynamic displacements and associated equivalent static forces.
f jo = Γ m j ψ j A (6.37)
Also, as discussed in Section 3.3, with these forces we can conduct a static structural analysis to
determine element forces (bending moment, shear force, and axial force) and stresses needed for
design of structural elements; no additional dynamic analysis is necessary. Figure 6.3 shows these
equivalent static forces along with the base shear and overturning moment. The internal story shear
force, Vmax i, and internal story moment, Mmax i, at an arbitrary level i can be obtained by applying
static equilibrium to the right-hand-side free-body diagram,
Vmax i = ∑f
j =i
jo
(6.38)
N
M max i = ∑ (h − h ) f
j =i
j i jo
(6.39)
Setting i equal to 1 in Equations 6.38 and 6.39 results in the base shear force, Vmax b, and overturn-
ing moment, Mmax b,
N
Vmax b = ∑f
j =1
jo
(6.40)
N
M max b = ∑h f
j =1
j oj
(6.41)
In the derivation of Equations 6.39 through 6.41, it is assumed that the floor weights are directly
in the center of the building frame; thus, making no contribution to the moment equilibrium
Generalized SDOF System Analysis 159
e quation. However, for buildings with off-center weights, a full moment static equilibrium analysis
should be conducted to account for the contribution of these off-center forces.
EXAMPLE 6.1
Given the three-story building shown in Figure E6.1 frame with damping of 5% and subjected to
a ground acceleration due to a 0.25g peak ground acceleration (PGA) earthquake characterized
by an 84.1% design spectrum, determine (i) peak displacements, (ii) maximum base shear, and
(iii) maximum floor overturning moments. Let us assume that beams are rigid. Each story has a
stiffness k = 326.3 kips/in and a shape function appropriate for a mid-rise building is to be used.
Solution
1. Determine the shape vector, ψj. This can be obtained using Figure 6.2 for a mid-rise
building function, ψj = x/H, where H = 36 ft.
1/ 3
ψ j = 2/ 3
1
2. Determine the generalized properties. The building frame can be modeled as a gen-
eralized SDOF system assuming that the beams are rigid and only lateral deformations
of columns occur. The stiffness of each floor is then obtained by summing the column
lateral stiffnesses at each level. The generalized mass, stiffness, and force of the general-
ized SDOF system are calculated as follows:
∑
3
= 1.5 m 19/18 m 1/3 k
j =1
Substituting the given values for mass and stiffness, the following properties of the
generalized SDOF are calculated:
Generalized mass,
3
1 1
k = ∑ k ⋅ (∆ψ )
j =1
j j
2
=
3
k = 326.3 kips /in(1000 lb /kip) = 108,767 lb /in
3
160 Introduction to Earthquake Engineering
Generalized force,
3
W 100 kips(1000 lb /kip)
L = ∑ m ⋅ ψ = 1 . 5 m = 1 . 5 g = 1 .5
j =1
j j
386.4 in/s2
= 388.2 lb s2 /in
3. Determine the natural period, frequency, and participation factor of the generalized
SDOF system:
Natural period,
m 273.2 lb s2 /in
Tn = 2π = 2π = 0.315 s
k 108,767 lb /in
Natural frequency,
k 108,767 lb /in
ωn = = = 19.95rad /s
m 273.2 lb s2 /in
Participation factor,
L 388.2 lb s2 /in
Γ = = = 1.421
m 273.2 lb s2 /in
where:
η is the PGA scale for a given earthquake; in this case 0.25
A 261.8 in /s2
D= = = 0.658 in
ωn2 (19.95 rad /s)2
We can now determine the maximum floor displacements using Equation 6.35,
5. Determine the equivalent static story forces: Determine the equivalent static story forces
using Equation 6.37,
f jo = Γ m j ψ j A
Generalized SDOF System Analysis 161
50 kips
k 100 kips 12 ft
k 12 ft
100 kips
12 ft
k
6. Determine base shear and overturning moment: Determine the base shear force using
Equation 6.40,
Vmax b = ∑f j =1
jo = 32.1kips + 64.2 kips + 48.1kips = 144.4 kips
Mmax b = ∑h f
j=1
j oj = 32.1kips(12 ft ) + 64.2 kips( 24 ft ) + 48.1kips(36 ft ) = 3658 kip ft
We can also obtain the internal story shear forces and moments at each level using
Equations 6.38 and 6.39, respectively. Alternatively, we can draw shear force and
bending moment diagrams by treating the building as a cantilever beam in order to
obtain the internal story shear forces and moments shown in Figure E6.2. This diagram
also summarizes the results of the analysis.
64.2 kips 12 ft
0.62 in
48 –578
12 ft
32.1 kips 0.31 in
112 –1925
12 ft
FIGURE E6.2 Lateral forces, internal shear, and internal moment diagram.
162 Introduction to Earthquake Engineering
6.2.1 Equation of Motion
The equation of motion for a tower subjected to a time-dependent input force or ground excita-
tion can be formulated in terms of z(t) only, provided we can find an appropriate shape of the
deformed system as it vibrates, ψ(x). The generalized equation of motion in terms of z(t) can again
be derived using the principle of virtual work as described in the last section; the resulting second-
order ordinary differential equation is of the same form as Equation 6.29 and is rewritten below for
convenience,
+ cz
mz = −Lu
+ kz g (t )
In the continuous system, integration is used to determine properties of the generalized
SDOF as follows:
L
m =
∫ 0
m( x )[ψ( x )]2 dx is the generalized mass
L
k =
∫ 0
EI ( x )[ψ ′′( x )]2 dx is the generalized stiffness
L
L =
∫ 0
m( x )ψ( x ) dx is the generalized force
Generalized damping, c , is not explicitly determined; it’s estimated using the damping ratio,
L is the tower height,
δz z(t)
ut
z(t)
EI(x), m(x)
δu u(x,t) = ψ(x)⋅z(t)
~
m
L
x
~
k
ug(t) ug(t)
FIGURE 6.4 Tower with distributed properties and equivalent generalized SDOF system.
Generalized SDOF System Analysis 163
Again, this generalized equation of motion is the same as the one that characterizes the vibration
of the SDOF oscillator introduced in Chapter 3; thus, all analysis procedures described so far can
be used to determine the response z(t). With z(t) and the shape function ψ(x), we can determine a
function of the shape into which the tower vibrates, which is the relative displacement at an arbitrary
height x,
u( x, t ) = ψ ( x )z(t ) (6.42)
The accuracy of the results in this case depends on how close the assumed shape function
tracks the actual shape of the deformed tower. A satisfactory approximation for the shape func-
tion, ψ(x), must satisfy the geometric boundary conditions of the tower, which for this case
include zero ground lateral relative displacement and rotation; that is, u(0) = 0 and u′(0) = 0,
where the prime notation indicates differentiation with respect to x, u′(x) = du/dx. Also, since z
is only a function of t, these boundary conditions reduce to ψ(0) = 0 and ψ′(0) = 0. Furthermore,
the function ψ(x) must be normalized at the top of the tower such that ψ(L) = 1.0, where L is the
height of the tower.
This analysis can be used to determine internal stress resultants in the tower, which can then
be used to determine the stresses. The analysis procedure can be used with a variety of cantilever
systems subjected to different time-dependent loads (Chopra 2012). Like the shear building, here
we assume a tower subjected to a time-dependent ground lateral displacement of ug(t) as shown in
Figure 6.4, which gives a total lateral displacement at a height x of
ut ( x, t ) = u( x, t ) + ug (t ) (6.43)
where:
k
ωn = is the generalized natural frequency
m
L
Γ = is the generalized participation factor
m
ζ is an estimate of the damping ratio
m
generalized natural period is given as before, Tn = 2π
k
design response spectrum. The maximum displacement at the top of the tower based on response
spectrum analysis is
z = Γ ⋅ D (6.44)
where:
D is the spectral displacement
The maximum dynamic displacement relative to the ground along the entire tower can be
obtained using this displacement
uo ( x ) = z ⋅ ψ( x ) = Γ ⋅ D ⋅ ψ( x ) (6.45)
A
uo ( x ) = Γ ⋅ 2 ⋅ ψ( x ) (6.46)
ωn
fo ( x, t ) = m( x ) ⋅ ψ( x ) ⋅ Γ ⋅ A (6.47)
fo ( x, t ) = ωn2 ⋅ m( x ) ⋅ ψ( x ) ⋅ Γ ⋅ D (6.48)
The internal shear force, Vmax, and internal story moment, Mmax, at an arbitrary height, x, can be
obtained by applying static equilibrium,
L L
Vmax ( x ) =
∫ fo (ξ )d ξ = Γ A
∫ m(ξ)ψ(ξ)dξ (6.49)
x x
L L
M max ( x ) =
∫ (ξ − x ) fo (ξ )d ξ = Γ A
∫ (ξ − x)m(ξ)ψ(ξ)dξ (6.50)
x x
where:
ξ ranges from x to L
Setting x equal to 0 in Equations 6.49 and 6.50 gives the base shear force, Vmax b, and overturning
moment, Mmax b,
M max b = Γ A
∫ xm( x)ψ( x)dx (6.52)
0
EXAMPLE 6.2
The reinforced concrete chimney shown in Figure E6.3 is subjected to a ground acceleration due
to an earthquake characterized by an 84.1% design spectrum, scaled to 0.25g PGA. Determine
(i) period, (ii) peak displacement, (iii) maximum base shear, and (iv) maximum overturning moment
of the chimney. Assume diameter, d = 3.0 ft; thickness = 4 in; modulus of elasticity, Ec = 3600 ksi;
concrete weight, γc = 150 pcf; damping, ζ = 5%; and the shape function shown in Figure E6.3.
Solution
1. Determine the properties of the chimney: The properties per unit length at an arbitrary
height, x, are as follows:
Cross-sectional area at x,
m(x ) = Area(x ) ⋅ γ c /g = 2.79 ft 2(150 pcf ) / 32.2 ft /s2 = 13.0(lbs ⋅ s2 /ft)) /ft
Moment of inertia at x,
Flexural stiffness at x
EcI(x ) = 3600 ksi(1000 lb /kip)(12 in /ft )2( 2.52 ft 4 ) = 1.307 ×109 lbs ⋅ ft 2
2. Determine the generalized properties: The generalized mass, stiffness, and force of the
generalized SDOF system are calculated as follows:
Generalized mass,
L 30 ft
x2 2
=
m
∫ 2
m( x )[ψ( x )] dx = 13 lb s /ft 2 2
∫ 2
(30 ft )2 dx = 78.1lb s /ft
0 0
Generalized stiffness,
d 2ψ( x ) 2
ψ ′′( x ) = =
dx 2 (30 ft )2
L 30 ft 2
2
k =
∫ EI( x )[ψ ′′( x )] dx = 1.307 ×10 lbs ⋅ ft
2 9 2
∫
5
(30 ft )2 dx = 1.94 ×10 lb /ft
0 0
166 Introduction to Earthquake Engineering
L 30 ft
13lb s2 /ft 2
L =
∫ m( x )ψ( x )dx =
(30 ft )2 ∫ x dx = 130 lb s /ft
2 2
0 0
3. Determine the natural period, frequency, and participation factor of the generalized
SDOF system.
Natural period,
m 78.1lb − s2 /ft
Tn = 2π = 2π = 0.126 s
k 1.94 ×105 lb /ft
Natural frequency,
Participation factor,
L 130 lb − s2 /ft
Γ = = = 1.667
m 78.1lb − s2 /ft
4. Determine maximum displacement at the top of the tower: This displacement is z(t) or
D due to 84.1% design spectrum, scaled to 0.25g PGA; first, enter the response spec-
trum in Figure 5.11 with Tn = 0.126 s, which is greater than 1/8 s, so the acceleration is
constant,
where:
η is the PGA scale for a given earthquake; 0.25 in this case
x2
ψ(x) =
L2
30 ft
A 261.8 in/ s2
D= 2
= = 0.106 in
ωn ( 49.8 rad /s)2
We can now determine the maximum displacement at the top of the tower,
5. Determine base shear and overturning moment: Determine the base shear force using
Equation 6.51,
Vmax b = Γ AL = 1.667( 261.8 in /s2 )(130 lb s2 /ft )(ft /12 in)(kip /1000 lb) = 4.73 kips
L 30 ft
13 lb s2 /ft 2
Mmax b
∫
= Γ A xm( x )ψ( x )dx = 1.667(21.8 ft /s )2
(30 ft )2 ∫ x dx = 106 kip ⋅ ft
3
0 0
PROBLEMS
6.1 Solve Example 6.1 with a shape function given for a low-rise building.
6.2 Solve Example 6.1 with a shape function of ψ(x) = (x/H)2, where H is the building
height.
6.3 Use the generalized SDOF analysis for the following building frame to determine the
story displacements, story forces, base shear, and overturning moment due to an earth-
quake characterized by an 84.1% design spectrum, scaled to 0.25g PGA. Let us assume
damping of 5% and a shape function of ψ(x) = sin (πx/2H), where H is the building
height.
w2 = 150 kips
k1 = 40 kips/in 15 ft
6.4 Use the generalized SDOF analysis for the following building frame to determine the
story displacements, story forces, base shear, and overturning moment due to an earth-
quake characterized by an 84.1% design spectrum, scaled to 0.25g PGA. Let us assume
damping of 5% and a shape function given for a low-rise building.
w2 = 20 kips
k2 = 10 kips/in 12 ft
w1 = 50 kips
k1 = 20 kips/in 14 ft
20 ft 20 ft
168 Introduction to Earthquake Engineering
6.5 Given a five-story building frame subjected to a ground acceleration due to a 0.25g PGA
earthquake characterized by the 84.1% design spectrum, use generalized SDOF analy-
sis to determine (i) peak displacements, (ii) maximum base shear, and (iii) m aximum
floor overturning moments. k = 326.3 kips/in, w5 = 50 kips, and the weights on the
other four floors equal 100 kips. Let us assume rigid beams, damping of 5%, and a shape
function given for a mid-rise building.
6.6 Solve Problem 6.5 with a shape function given for a high-rise building.
6.7 Solve Example 6.2 with a shape function of ψ(x) = 1.5 (x/L)2 – 0.5 (x/L)3.
6.8 Solve Example 6.2 with a shape function of ψ(x) = 1 – cos(πx/2L).
REFERENCES
Chopra, A. K., Dynamics of Structures: Theory and Applications to Earthquake Engineering, 4th edition,
Prentice-Hall, Upper Saddle River, NJ, 2012.
Naeim, F., The Seismic Design Handbook, 2nd edition, Springer, Dordrecht, The Netherlands, 2001.
7 Multi-Degree-of-Freedom
System Analysis
After reading this chapter, you will be able to:
169
170 Introduction to Earthquake Engineering
P(t)
And so on
h
The eigen matrix operation can be viewed as a linear transformation that maps vectors into mul-
tiples of themselves. That is,
Ax = λx or [ A]{ x} = λ{ x} (7.1)
where:
I is the identify matrix
The nonzero solutions to this equation require that λ be chosen so that
The values of λ that satisfy this relationship correspond to eigenvalues of matrix A; each λ yields
an eigenvector {x} solution of Equation 7.1 to within a multiplicative constant. That is, scaling the
eigenvector by any constant factor will still satisfy Equation 7.1. Consequently, each eigenvector is
normalized with respect to one of its elements, usually the first element.
m4
u4
4
k4
m3
3 u3
k3
m2 u2
2
k2
m1
1 u1
k1
FIGURE 7.2 Idealized four-story shear building system and mode shapes.
Multi-Degree-of-Freedom System Analysis 171
These solutions are practically intractable via hand calculations for systems with more than two
degrees of freedom, which is why solutions are regularly found via computer using one of the widely
available numerical packages, such as MATLAB®. With this program, we can use the eig operator
to determine the eigenvectors denoted as phi and the eigenvalues denoted as lambda as follows,
In the sections that follow, we present the application of the eigenvalue problem to the solution of
structural dynamics problems. This solution will then be extended to encompass the calculation of
the maximum shear and moment response of MDOF systems.
u2
m2 u 2 – u1 m2u2
k2 (u2 – u1)
k2 (u2 – u1)
k2
k2
u1
k2 (u2 – u1)
m1 m1u1
k2 (u2 – u1)
k1u1
k1u1
k1
k1
k1u1
u1
where m1 is the mass of the first floor, which can be determined following the procedure discussed
in Section 3.1.2, and k1 and k2 are the story stiffnesses of the first and second levels, respectively,
which can be determined by adding the lateral stiffnesses of all columns in the story. Each story can
be treated as a portal frame as discussed in Chapter 3; thus, for a story of height h, column flexural
stiffness EI, and assuming rigid floor diaphragms (rigid frame beams), the lateral stiffness of a col-
umn is 12EI/h3 and the stiffness of story j is kj = ∑columns 12EI/h3.
Equilibrium of mass 2:
In general,
where [m] is the mass matrix, which is a diagonal matrix (nonzero elements only along the forward
diagonal), and [k] is the stiffness matrix; this can alternatively be obtained by directly finding each
coefficient kij, which represents a force needed at level i to hold a unit displacement at level j, holding
zero displacements at all other levels.
Notice that an Nth-degree-of-freedom system results in N-dependent equations, or N × N mass
and stiffness matrices.
where:
ωn = k m is the natural circular frequency with units of radians per second (rad/s), which can
be related to the period as Tn = 2π /ωn
Taking the first and second derivatives of Equation 7.9 with respect to time,
This equation is similar to Equation 7.2 and has a nontrivial solution (one that gives nonzero
values for {u}) if Equation 7.3 is satisfied. That is,
This is the characteristic (or eigen) equation of the system, also known as the frequency equa-
tion because expanding the determinate results in an Nth degree polynomial in terms of ωn2 . That
is, Equation 7.14, in general, should be satisfied for N values of ωn2 . We can then solve the homoge-
neous system of equations (Equation 7.12) for a1, a2, … , aN for each value of ωn2 , totaling N modes
of vibration.
Solving this eigenvalue problem is convenient using a computer program such as MATLAB; for
this case, the eig operator includes the stiffness, [k] and the mass, [m] matrices to determine eigen-
vectors, denoted by phi and eigenvalues denoted by lambda, as
For illustration purposes, let us conduct the analysis of the two-degree-of-freedom case in Figure
7.3 using hand calculations. Recall that the assumed displacement response vector is given as
Taking two derivatives of Equation 7.16 with respect to time and substituting into Equation 7.7
gives the following equation:
k + k2 − m1ωn2 −k2
det 1 =0 (7.19)
−k2 k2 − m2ωn2
174 Introduction to Earthquake Engineering
Expanding this determinate results in a second-order polynomial in ωn2 , the frequency equation,
We use the quadratic formula, x = (−b ± b2 − 4ac ) / 2a , to find the roots of Equation 7.21,
which are the eigen (characteristic) values:
or
3− 5 k 3+ 5 k
ωn21 = and ωn22 = (7.23)
2 m 2 m
k k
ωn1 = 0.618 and ωn 2 = 1.618 (7.24)
m m
Frequencies are usually listed in ascending order; that is, from the smallest to largest.
Also, each of these eigenvalues is associated with a distinct eigenvector (or mode shape). The
actual values of the eigenvector elements in Equation 7.18, a1 and a2 for this case, are indeterminate.
However, each eigenvector can be obtained by setting the values of the top element equal to one,
2
and solving for the values of the other eigenvector elements relative to top value. Substituting ωn1
into Equation 7.18 yields
2k − m 3 − 5 k −k
2 m
a1 0
= (7.25)
3 − 5 k a2 0
−k k − m
2 m
1 1
m2
k2
a1 0.618 a1 –1.618
φ1 = = φ2 = =
0.618 a2 1
1 a2 2
1
–1.618
m1
k k
ω 1 = 0.618 m ω 2 = 1.618 m
k1
These eigenvectors are referred to as normal modes because they are normalized so that the
displacements of the mass at the top of the structure are equal to one. The resulting eigenvectors
describe the mode shapes and are defined as {φ}j (this was the process followed in Chapter 6 to
obtain a shape vector that describes the deflected shape of a vibrating shear building). A summary
of these results, including the normal modes and frequencies, is shown in Figure 7.4.
Regardless of how the eigenvectors are normalized, the resulting mode shapes are orthogonal;
that is, for two distinct mode shapes of the structure, r ≠ s,
u1 = φ11q1 + φ12q2
(7.31)
u2 = φ21q1 + φ22q2
where we now assume a new set of harmonic displacement functions with constants Aj and Bj,
The matrix [φ] is a coordinate transformation matrix (also known as the modal matrix) that maps
the original coordinates {u} to the principal coordinates {q}. The inverse transformation is
The orthogonality condition given by Equation 7.29 implies that [φ]−1 = [φ]T. Thus, we can then
rewrite Equation 7.34 as
Substituting the transformation relationship, Equation 7.33, into the equations of motion,
Equation 7.8, and premultiplying by [φ]T yields
The matrices [M] and [K] are diagonal because of the orthogonality condition given by
Equation 7.30 (i.e., all off diagonal elements are zero). These are known as modal mass and modal
stiffness matrices, respectively. The diagonal elements of these matrices are given by
N N
Notice that Equations 7.36 are now uncoupled, or independent. That is,
Mi qi + K i qi = 0 (7.38)
If each equation is divided by the associated modal mass, the equations can be rewritten as
So, each equation has the same form as the equation of motion for the SDOF system we described
in detail in Chapter 3. This implies that we can use the solutions to the different loadings or forcing
functions examined previously, including response spectrum analysis.
To illustrate the process of decoupling the equations of motion, let us revisit the two-degrees-of-
freedom case introduced earlier in this chapter. Using Equation 7.36, we get the mass transforma-
tion coefficients M1 and M2:
This also proves the orthogonality condition ({φ1}T [m]{φ2} = {φ2}T [m]{φ1} = {0} and {φ1}T [k]
{φ2} = {φ2}T [k]{φ1} = {0}) since the off-diagonal elements of the two matrices are zero. That is,
m 0 −1.618 −1.618
{φ1}T [m]{φ 2 } = [0.618 1] = [0.618m m] =0 (7.42)
0 m 1 1
EXAMPLE 7.1
For the two-story building frame shown in Figure E7.1, determine (1) mass and stiffness matrices,
(2) periods and modal matrix (normalized mode shapes), and (3) the stiffness and modal matrices.
Assume beams are rigid. k = 20 kips/in and m = 0.4 kips · s2/in.
Solution
1. Determine the mass and stiffness matrices. These can be obtained following the process
shown in Figure 7.3 by setting m1 = m2 = m, k1 = 2k, and k 2 = k, and applying equilib-
rium to the free-body diagrams of each mass, resulting in equations similar to Equations
7.5 and 7.6. The two equations can be combined into a matrix equation as shown in
Equation 7.7:
m 0 0 .4 0
[ m] = = kip ⋅ s2 /in
0 m 0 0.4
k m 15 ft
2k 15 ft
3k −k 60 −20
[ k ] = = kips/in
−k k −20 20
2. Determine the periods and modal matrix. The natural frequencies and mode shapes
can be determined by solving for the eigenvalues and eigenvectors using a standard
eigen solution; natural frequencies can be obtained using the eigenvalues and the mode
shapes with the eigenvectors. Also, the periods are obtained using the natural frequen-
cies. The frequency equation (Equation 7.20) resulting from the expansion of the eigen
equation is
m2ωn4 − 4 kmωn 2 + 2k 2 = 0
Using the quadratic formula, we can find the roots of this equation, which are the
eigenvalues:
w n2 =
4km ± ( 4km)2 − 8m2k 2 4 ± 2 2 k
2m2
=
2 m
= 2± 2
20 kips /in
(
0.4 kip ⋅ s2 / in
)
The frequencies are (from smallest to largest)
(
3k − 2 − 2 k
) −k
a1 0
=
2.414
−1 a1 0
=
or
−1 0.414 1 0
−k k − 2 − 2 k a2 0
( )
0.414 −2.414
[φ] =
1 1
Multi-Degree-of-Freedom System Analysis 179
3. Determine the modal mass and stiffness matrices. Use Equations 7.37 to determine the
elements of the modal mass and stiffness matrices. Or, we can use the matrix operations
in Equation 7.36:
Modal mass matrix:
4. Alternatively, we can write a MATLAB script to perform all the operations; the most
important of the operations being the eig operator in Equation 7.15:
The results of this script are the same as those of parts 1–3:
omegas =
5.4120 0
0 13.0656
periods =
1.1610
0.4809
norm_phi =
0.4142 −2.4142
1.0000 1.0000
M =
0.4686 0.0000
0.0000 2.7314
K =
13.7258 0.0000
0 466.2742
180 Introduction to Earthquake Engineering
utj (t ) = u j (t ) + ug (t ) (7.44)
For MDOF systems, we can substitute the second derivative of Equation 7.44 into the first term
of Equation 7.8 to obtain the equations of motion including ground excitation,
Substituting modal mass and stiffness relations, Equations 7.37, into Equation 7.46, we get
where the damping is assumed to be constant for all mode shapes. For elastic analysis of MDOF
systems, where damping effects are significant, there are a number of approaches used to establish
the damping matrix, including a relationship based on the stiffness and mass matrices. This can be
found in conventional structural dynamics texts such as Chopra (2012) and Villaverde (2009). Also,
we can now define the participation factor for each mode shape (the modal participation factor) as
where Lj is the generalized force associated with the jth mode shape.
Multi-Degree-of-Freedom System Analysis 181
EXAMPLE 7.2
Consider the two-story building frame shown in Figure E7.1 and determine the participation factors.
Solution
1. Determine the participation factors. First, apply the summation form of Equation 7.50:
∑
2
∑
2
The results of this script are the same as those of part (i):
LT =
0.5657
-0.5657
MT =
0.4686
2.7314
par_fac =
1.2071
-0.2071
Also, in MDOF systems, a modal participation factor can be viewed as providing a measure
of the degree to which the jth mode contributes (or participates) to the total dynamic response.
Equation 7.50 is essentially the same as Equation 6.32 for the generalized SDOF case, except the
use of N eigenvectors rather than a single assumed mode shape. Therefore, we can use the formula-
tions presented in Chapter 6 to obtain the response of a system due to each eigenvector. That is, the
original system of N simultaneous differential equations of motion is now transformed into a system
182 Introduction to Earthquake Engineering
of N-independent differential equations, each of which can be solved using the SDOF procedures
described in Chapters 3–5. Thus, for linear elastic systems the total displacement can be obtained
using superposition of the modal contributions:
∑
N
u j (t ) = φ ji ⋅ qi (t ) or {u(t )} = {φ}1 q1(t ) + {φ}2 q2 (t ) + + {φ} N qN (t ) (7.51)
i=1
where i denotes the level, from 1 to N, and j the range of eigenvectors, from 1 to N. With these
displacements, we can compute equivalent lateral forces, which can then be used to conduct a
static structural analysis to determine the internal element forces and stresses. This superposition
approach produces the entire time history of the structural response (displacements, forces, etc.).
However, for design of various elements of the structure, we only need the absolute maximum val-
ues of the response. This is the basis for the code-based seismic response history modal analysis
procedure introduced in Section 8.7. A more efficient approach is to use the response spectrum
analysis described in the next section.
q jo = Γ j ⋅ D j (7.52)
These maximum displacements can be distributed to the other floors using the normalized eigen-
vectors, or mode shapes. That is, the maximum displacement value of the ith floor, induced, when
only the jth mode shape is excited, is
The displacement subscript ij can be interpreted as the contribution to the total response at level
i due to mode of vibration j. In matrix form, each matrix column represents the floor displacements
associated with each mode shape, that is,
Aj
uijo = φij ⋅ q jo = φij ⋅ Γ j ⋅ (7.55)
ωnj2
The maximum acceleration value of the ith floor, induced when only the jth mode shape is
excited, is
where:
[ωn2 ] is a diagonal matrix of the square modal frequencies (also known as the spectral matrix)
[A] is a diagonal matrix of spectral accelerations
The equivalent static forces associated with these floor accelerations can be obtained using
Newton’s second law; the product of the associated story mass and maximum story acceleration;
that is, the inertial force at level i due to vibration mode j (using Equations 7.56) is
From static analysis, we can determine the force at the base of the building, which is known as
the base shear, Vjb, for each mode j,
N
V jb = f1 j + f2 j + f3 j + = Γ j ωnj2 D j ∑ miφij = Γ j ωnj 2 D j L j or V jb = Γ j A j L j (7.60)
i =1
{Vb } = [ f ]T {1} = [Γ][ A][φ]T [m]{1} = [Γ][ D] ωn2 [φ]T [m]{1} (7.61)
Figure 7.5 shows the equivalent static forces along with the base shear and overturning moment
for the jth mode shape. To obtain the overturning moment, we can apply static equilibrium to the
free-body diagram.
184 Introduction to Earthquake Engineering
mN qj
fNjo uNjo
. kN
.
.
fijo mi uijo
. ki
.
.
.
. hi
.
m1 u1jo
f1jo
k1
Mjb
Vjb
FIGURE 7.5 Maximum dynamic displacements and associated equivalent static forces.
The inertial forces (or lateral story forces) at level i due to vibration mode j can also be written
in terms of the base shear by substituting the second of Equations 7.60 as Γ j A j = V jb /L j into the
second of Equations 7.58:
V jb m jφijV jb (7.62)
fij = miφij Γ j A j = miφij or fij =
Lj ∑ Nj=1 m jφ ji
Also, writing the second of Equations 7.60 in the form of Newton’s second law (F = mA), we get
V jb = M ej A j (7.63)
2
e
M = Γ jLj or M =eL2j
=
( ∑ iN=1 miφij ) (7.64)
j j
Mj ∑ iN=1 miφij2
Substituting into Equation 7.63 the definition of the mass in terms of the floor weights, mi = wi/g
(weight associated with the ith floor divided by the acceleration due to gravity),
V jb =
( ∑ iN=1 wiφij ) Aj A
= W je j
(7.65)
∑ iN=1 wiφij2 g g
W e
=
( ∑ iN=1 wiφij ) (7.66)
j N 2
∑ i=1 wφ
i ij
Multi-Degree-of-Freedom System Analysis 185
Combining the contribution of all the modes of vibration, we get the total weight of the
building as
N
Wtotal = ∑W
j =1
j
e
(7.67)
This implies that the effective weight of the jth mode shape is the fraction of the total weight that
participates in the jth mode of vibration. The code (ASCE-7) requires including enough modes of
vibration in the analysis to obtain an effective weight of at least 90% of the actual building weight,
as will be discussed in Section 8.6.
Displacement, acceleration, and lateral story forces resulting from different mode shapes must be
combined to acquire the total response, though a simple summation is not used since modal maxima
generally occur at different times during the response history. Furthermore, response spectra only
provide the values of the modal maxima and not the time at which each value occurs. Therefore,
appropriate combination rules have been developed to obtain the total response. These rules are
based on random vibration theory to estimate the average maximum response. The most popular
rules include the SRSS and CQC methods, the latter being more complicated, but yielding more
accurate results for a wider range of mode shapes. The following relationships for the SRSS and CQC
are given in terms of generic maximum modal values, Rj, which are intended to represent maximum
values due to vibration mode j for any response parameter, such as displacement, acceleration, etc.
The total response for the SRSS rule
(7.68)
Rmax ≈ ∑ Nj =1 R 2j
(7.69)
Rmax ≈ ∑ iN=1 ∑ Nj =1 R jρij R j
where:
8ς 2 (1 + βij )βij1.5
ρij = (7.70)
(1 + βij2 )2 + 4βijς 2 (1 + βij )2
ρij is the cross-modal coefficient that varies from 0 to 1 (1 for the case of i = j), and is generally
expressed in terms of the modal frequencies of two distinct mode shapes, βij = ωj/ωi and d amping
characteristics, which reduces to the damping ratio given by Equation 3.32 (repeated here for con-
venience, ζ = c/ccr) when the modal damping is constant for the entire modal spectrum. Also, when
the modal frequencies are well separated, this matrix tends to the identity matrix and the CQC rule
approaches the SRSS rule results.
To determine an upper bound to both of these rules, we can sum the maximum absolute values as
N
Rmax ≤ ∑R
j=1
j (7.71)
Also, as discussed in Section 3.3, with the maximum dynamic displacement or acceleration, we can
conduct a static structural analysis to determine element forces (bending moment, shear force, and axial
force), and stresses needed for design; no additional dynamic analysis is necessary. These maximum
element force or stress results from each mode shape can be combined using either SRSS or CQC.
186 Introduction to Earthquake Engineering
EXAMPLE 7.3
Consider the two-story building frame of Figure E7.1 and determine (1) effective masses, (2) peak
displacements, (3) maximum equivalent static floor forces, (4) maximum base shear, and (5) maxi-
mum overturning moment. Assume beams are rigid, damping of 5%, and that the frame is sub-
jected to a ground acceleration due to an earthquake characterized by the 84.1% design spectrum
scaled to 0.25g PGA, the solution is given as follows.
Solution
1. Determine the effective masses. Use Equation 7.64:
M1e =
L12
=
(∑i2=1 miφi1) = (m1φ11 + m2φ21)2 = (0.4(0.414) + 0.4(1))2 = 0.320 = 0.683
M1 ∑ i2=1 miφi21 2
m1φ11 2
+ m2φ21 0.4(0.414)2 + 0.4(1)2 0.469
2
L2
M = 2 =
e (∑i2=1 miφi1) = (m1φ11 + m2φ21)2 = (0.4(−2.414) + 0.4(1))2 = 0.320 = 0.117
2
M2 ∑ i2=1 miφi21 2
m1φ11 2
+ m2φ21 0.4(−2.414)2 + 0.4(1)2 2.731
The operation can be performed using MATLAB after determining participation fac-
tors, using the same factors, MT and LT; that is,
Eff_mass = LT.∧2./MT % .∧ operation squares each element of LT vector
Tn1 = 1.16 s, which is greater than 0.66 s, so the acceleration is given as,
A1 = η · 1.8g/Tn1 = 0.25(1.8 s)(386.4 in/s2)/1.16 s = 149.9 in/s2, and
Tn2 = 0.48 s, which is greater than 1/8, but less than 0.66 s, so the acceleration is
constant,
A 2 = η · 2.71g = 0.25(2.71)(386.4 in/s2) = 261.8 in/s2,
where:
η is the PGA scale factor for a given earthquake, in this case 0.25
The maximum spectral displacements are
A1 149.9 in/s2
D1 = = = 5.12 in
ωn21 (5.41rad/s)2
A1 261.8 in/s2
D2 = = = 1.52 in
ωn 2 (13.1rad/s)2
2
Now, determine the maximum floor displacements using Equation 7.53, uijo = φij ⋅ Γ j ⋅ D j
0.414 2.56
ui1o = φi1Γ1 ⋅ D1 = (1.207)(5.12 in) = in
1 6.18
Finally, combine these displacements using the SRSS rule (Equation 7.68) to obtain the
peak displacements at each level:
[f ] = [m][φ][Γ][ A]
0 .4 0 0.414 −2.414 1.2077 0 149.9 0
=
0 0 . 4 1 1 0 − 0 . 207 0 261 .8
0.4(0.414)(1.207)(149.9) 0.4(−2.414)(−0.207)( 261.8)
=
0.4(1.207)(149.9) 0.4(−0.207)( 261.8)
30.0 52.3
= kips
72.4 −21.7
Finally, combine these forces using the SRSS rule (Equation 7.68) to obtain the maxi-
mum floor forces at each level:
5. Determine maximum overturning moment. Use static equilibrium to obtain the contribu-
tion from each mode to the maximum overturning moment:
Finally, combine these using the SRSS rule (Equation 7.68) to obtain the maximum
overturning moment:
1. Determine the mass matrix, [m] from the given floor weights.
2. Determine the stiffness matrix, [k] from the column properties for a shear building.
3. With the stiffness and mass matrices, solve the associated eigenvalue problem for eigenval-
ues, ω2, which are used to determine the natural frequencies, ωn and periods, Tn.
4. After selecting an appropriate damping ratio, determine spectral accelerations (or dis-
placements) using the ordinates of the acceleration (or displacement) response spectrum of
the excitation for each of the natural periods.
5. With the eigenvalues, we can also determine the eigenvectors, which are normalized to
obtain the modal matrix, [φ].
6. Obtain the modal participation factors, {Γ} = [φ]T [m]{1}/[φ]T [m][φ]; (the contribution of
each mode shape to the total response).
7. Determine the displacements associated with each mode shape, [u] = [φ][Γ][D] = [φ][Γ]
[A][ω2]−1
where:
[Γ] = diagonal matrix of participation factors
[D] = diagonal matrix of spectral displacements
[A] = diagonal matrix of spectral accelerations
[ω2] = diagonal matrix of squared modal frequencies
8. The resultant maximum displacement at each node is obtained using the SRSS rule
(Equation 7.68) for each row vector: umaxi = (Σui2)½.
9. The matrix of lateral forces at each node is: [f] = [k][u].
10. The resultant maximum lateral force at each node is obtained using SRSS of each row vec-
tor: fmaxi = (Σfi2)½.
11. The column vector of total base shear forces is: {Vb} = [f]T{1}
12. The maximum base shear force is obtained using SRSS as: Vbmax = (ΣVi2)½.
13. The maximum overturning moment is obtained from SRSS using static equilibrium.
EXAMPLE 7.4
Given the three-story building frame shown in Figure E7.2 with damping of 5% and subjected to a
ground acceleration due to an earthquake characterized by the 84.1% design spectrum scaled to
0.25g PGA, determine (1) periods, (2) mode shapes, (3) peak displacements, (4) maximum equiva-
lent static floor forces, (5) maximum base shear, and (6) maximum overturning moment. Assume
beams are rigid. Each story has a stiffness k = 326.3 kips/in (Figure E7.2).
u3
50 kips
m3
k 12 ft u2 k3
100 kips
m2
k 12 ft
u1
100 kips k2
m1
12 ft
k
k1
FIGURE E7.2 Building frame schematic (left) and idealized MDOF structural model (right).
Multi-Degree-of-Freedom System Analysis 189
Solution
1. Determine the equations of motion, frequencies, periods, and shape vectors. The equa-
tions of motion for this case can be determined using D’Alembert’s principle by apply-
ing horizontal equilibrium to the free-body diagrams of the three masses shown in
Figure E7.3. The three equations of motion are
These three dependent equations of motion are second-order, linear, and homogeneous
differential equations with constant coefficients, and can be rewritten in the matrix form as
u3 – u2
m2u3
k3(u3 – u2) k3(u3 – u2)
k3
u2 – u1 k3(u3 – u2)
k3(u3 – u2)
k2(u2 – u1) m2u2
k2(u2 – u1)
k2
k2(u2 – u1)
k2(u2 – u1)
m1u1
k1u1 k1u1
k1
k1u1
u1
FIGURE E7.3 FBDs for each mass used to determine equations of motion.
190 Introduction to Earthquake Engineering
omegas =
18.3803 0 0
0 50.2160 0
0 0 68.5963
periods =
0.3418
0.1251
0.0916
norm_phi =
0.5000 −1.0000 0.5000
0.8660 0.0000 −0.8660
1.0000 1.0000 1.0000
2. With the eigenvalues and eigenvectors, first we follow the same procedure as for the
generalized SDOF case using the first mode (following the same process as in Example
6.1). Then, we use matrix operations to complete the entire solution.
a. Use the first mode, φi1, as the shape vector:
0.5
φi1 = 0.866
1
b. Determine the generalized properties. The building frame can be modeled as a gen-
eralized SDOF system assuming that the beams are rigid and only lateral deforma-
tions of columns occur. The stiffness of each floor is then obtained by summing the
column lateral stiffnesses at each level. The generalized mass, stiffness, and force of
the generalized SDOF system are calculated as follows:
Substituting the given values for mass and stiffness, the following properties of the
“generalized SDOF,” which are the first elements of the modal mass and stiffness
matrices, are calculated as follows.
Multi-Degree-of-Freedom System Analysis 191
Generalized mass:
Generalized stiffness:
3
K1 = ∑ ki ⋅ (Δφi1)2 = 0.402 k = 0.402(326.3 kips/in)(1000 lb/kip) = 131144
, lb/in
i =1
Generalized force:
c. Determine the natural period, frequency, and participation factor for the first mode.
Natural period, which is the same as the MATLAB result in step 1:
M1 388.2lb-s2 /in
Tn1 = 2π = 2π = 0.3418 s
K1 ,
131144 lb/in
K1 131144
, lb/in
ωn1 = = = 18.38 rad/s
M1 388.2lb-s2 /in
Participation factor:
L1 482.9 lb-s2 / in
Γ1 = = = 1.244
M1 388.2lb-s2 / in
A1 261.8 in/s2
D1 = = = 0.775in
ω n21 (18.38 rad/s)2
We can now determine the maximum floor displacements using Equation 6.35:
0.5 0.482
n) = 0.835 in
uio = φi1Γ1 ⋅ D1 = 0.866 (1.244)(0.775in
1 0.964
192 Introduction to Earthquake Engineering
0.311
vs. Example problem 1 in Chapter 6, where u jo = 0.623 in
0.934
e. Determine the equivalent static story forces. Determine the equivalent static story forces
using Equation 6.37,
fio = Γ1miφi1A1
f. Determine base shear and overturning moment. Determine the base shear force using
Equation 6.40,
3
Vmax b = ∑ fio = 42.1 kips + 73.0 kips + 42.1 kips = 157.3 kips
1=1
3
Mmax b = ∑ hj foj = 42.1 kips(12 ft ) + 73.0 kips(24 ft) + 42.1 kips(36 ft ) = 3775 kip ⋅ ft
j =1
As in Chapter 6, we can obtain the internal story shear force and moment at each
level by drawing shear force and bending moment diagrams considering the building as
a cantilever beam. The diagrams also compare these results to those of the analysis of
Example 6.1, which are given in parentheses (Figure E7.4).
At this point, we could follow steps 2-a to 2-f for each of the remaining mode shapes
and then apply one of the combination rules to obtain the total response. Alternatively,
we could use the matrix operation process listed in Section 7.3, beginning with step 6 to
perform the analysis of all the modes simultaneously, see next step.
3. Use the matrix operations listed in Section 7.3:
FIGURE E7.4 Lateral forces, internal shear, and internal moment diagram for mode 1.
Multi-Degree-of-Freedom System Analysis 193
a. Obtain the modal participation factors, {Γ} = [φ]T[m]{1}/[φ]T[m][φ]: First, get the
modal matrix, [M] = [φ]T[m][φ],
0 .5 0.866 1 100 0 0 0 .5 −1 0 .5
[M] = −1 0 1 0 100 0 0.866 0 −0.866 / 386.4 in/ss2
0 .5 −0.866 1 0 0 50 1 1 1
0.388 0 0
[M] = 0 0.388 0 lb-s2 /in
0 0 0.388
0
0.0347
0.0893
0 2.576
b. Determine the displacements associated with each mode shape, [u] = [φ][Γ][A]
[ω2]−1. First, assemble the participation factors and square of the frequencies into
diagonal matrices:
Diagonal matrix of participation factors:
1.244 0 0
[Γ] = 0 −0.333 0
0 0 0.0893
337.8 0 0
2
[ω ] = 0
n 2521.6 0 rad2 /s2
0 0 4705.5
To obtain the diagonal matrix of accelerations, we first enter the response spec-
trum in Figure 5.11 with the following three values for Tn:
– Tn1 = 0.342 s, which is greater than 1/8 s so the acceleration is constant,
A1 = η · 2.71g = 0.25(2.71)(386.4 in/s2) = 261.8 in/s2;
– Tn2 = 0.1251 s, which is also greater than 1/8 s so the acceleration is constant,
A 2 = η · 2.71g = 0.25(2.71)(386.4 in/s2) = 261.8 in/s2; and
– Tn3 = 0.0916 s, which is less than 1/8 s so the acceleration is given as,
A3 = η · 11.7(Tn3)0.704g = 0.25(11.7) (0.0916)0.704 (386.4 in/s2) = 210.1 in/s2.
194 Introduction to Earthquake Engineering
261.8 0 0
[ A] = 0 261.8 0 in/s2
0 0 210.1
0 .5 −1 0.5 1.244 0 0
[u] = 0.866 0 −0.866 0 −0.333 0
1 1 1 0 0 0.0893
261.8 0 0 3 3 7 .8 0 0 −1
* 0 261.8 0 0 2521.6 0
0 0 210.1 0 0 4705.5
0.482 0.0346 0.0020
= 0.835 0 −0.0035 in
0.964 −0.0346 0.0040
c. The resultant maximum displacement at each node is obtained from the SRSS rule
(Equation 7.68) of each row vector: umaxi = (Σui2)½:
73.0k 4.2k
FIGURE E7.5 Resulting lateral forces at each node for a given mode of vibration: (a) Mode 1,
(b) Mode 2, and (c) Mode 3.
f. The column vector of total base shear forces is: {Vb} = [f]T{1},
g. The maximum base shear force is obtained from SRSS as: Vbmax = (ΣVi2)½,
The maximum story shear forces can be obtained using the SRSS rule at each level for
all the mode shape story shears obtained in step 3-f:
12 ft
73.1 kips 0.83 in
44 –524
12 ft
47.9 kips 0.48 in
116 –1906
12 ft
FIGURE E7.6 Lateral forces, internal shears, and internal moment diagrams.
h. The maximum overturning moment is obtained from SRSS using static equilibrium. We
first obtain the story overturning moments using static equilibrium at each level, for each
mode shape:
42.136 ( )
( ) + 73( 24) + 42.112 3774
{M1} = ( ) + 73(12)
42.124 kips ⋅ ft = 1887 kips ⋅ ft
42.112
( ) 505
−11.3(36) + 22.6(12) −135
{M2} = −11.3( 24) kips ⋅ ft = −271 kips ⋅ ft
−11.3(12) −135
2.4(36) − 4.3( 24) + 2.4(12) 13
{M3} = 2.4(24) − 4.3(12) kips ⋅ ft = 7 kips ⋅ ft.
2.4(12) 29
The maximum story overturning moments can now be obtained using the SRSS rule:
(3774)2 + (135)2 + (13)2
3777
2
{Mmax } = (1887) + ( 271) + (7) kips ⋅ ft = 1906 kips ⋅ ft
2 2
(505)2 + (135)2 + ( 29)2 524
Figure E7.6 summarizes the results of the maximum response obtained:
4. We can write a MATLAB script to perform all the operation as follows:
clear all % clears any previously defined variables
clc % clears the screen
g = 386.4; % acceleration due to gravity in in/s2
pA = 0.25* g; % peak ground acceleration
m = [100 0 0; 0 100 0; 0 0 50]/386.4 % mass matrix in kip · s2/in
k = [2 −1 0; −1 2 −1; 0 −1 1]*326.3 % stiffness matrix in kips/in
[phi,lam]=eig(k,m); % compute eigenvalues, lam and eigenvectors, phi
omegas=sqrt(lam) % determine and show frequencies from eigenvalues
periods = 2*pi*diag(inv(omegas)) % determine the periods from omegas
% mode shapes by normalizing eigenvectors to get top displ equal to 1
[N_rows, N_cols] = size(phi); % finds N, the size of the matrices
for i = 1:N_cols; % loops over the N modes to normalize them
norm_phi(:,i) = phi(:,i)./phi(N_rows,i);
end
norm_phi % display normalized modal matrix
% The spectral accelerations from an applicable response spectrum
Multi-Degree-of-Freedom System Analysis 197
m=
0.2588 0 0
0 0.2588 0
0 0 0.1294
k=
652.6000 –326.3000 0
−326.3000 652.6000 −326.3000
0 –326.3000 326.3000
omegas =
18.3803 0 0
0 50.2160 0
0 0 68.5963
periods =
0.3418
0.1251
0.0916
norm_phi =
0.5000 −1.0000 0.5000
0.8660 0.0000 −0.8660
1.0000 1.0000 1.0000
SA =
261.7860 261.7860 210.0515
198 Introduction to Earthquake Engineering
SD =
0.7749 0.1038 0.0446
LT =
0.4829
−0.1294
0.0347
MT =
0.3882
0.3882
0.3882
par_fac =
1.2440
−0.3333
0.0893
ui_max =
0.4820 0.0346 0.0020
0.8348 –0.0000 −0.0035
0.9640 –0.0346 0.0040
u_maxsrss =
0.4832
0.8348
0.9646
f=
42.1411 22.5833 2.4277
72.9905 −0.0000 −4.2049
42.1411 −11.2917 2.4277
f_maxsrss =
47.8724
73.1115
43.6951
V=
157.2726 11.2917 0.6505
V_maxsrss =
157.6788
OTM =
1.0e+003 *
3.7745
−0.1355
0.0156
OMT_maxsrss =
3.7770e+003
Multi-Degree-of-Freedom System Analysis 199
PROBLEMS
7.1 Using the definition of stiffness and mass influence coefficients, formulate the equation of
motion for the frame shown below. Assume the beams are rigid.
m/2
p2(t)
h EI
m
p1(t)
h EI
2h
7.2 Formulate the equation of motion (in matrix form) using D’Alembert’s principle for the
given three-story building, which has rigid beams and flexible steel (E = 29,000 ksi) col-
umns with total moment of inertial for each floor as shown.
400 kips
I = 1240 in4 14 ft
7.3 Formulate the equation of motion (in matrix form) using D’Alembert’s principle for the
given two-story building, which has rigid beams and flexible steel (E = 29,000 ksi) col-
umns with total moment of inertial for each floor as shown.
w2 = 20 kips
20 ft 20 ft
200 Introduction to Earthquake Engineering
7.4 Formulate the equation of motion (in matrix form) using D’Alembert’s principle for the
given three-story building, which has rigid beams and flexible steel (E = 29,000 ksi) col-
umn sections as shown.
WDL = 1 kips/ft
W14×109 W14×120
12 ft W14×109
20 ft 20 ft
7.5 Formulate the equation of motion (in matrix form) using D’Alembert’s principle for the
given two-story building, and determine the natural frequencies, periods, normalized
modal matrix, and participation factors.
3m
7.6 The following two-story building is supported with steel frames spaced at 15 ft on center,
and has the floor load shown plus wall load of 20 psf. Formulate the equation of motion (in
matrix form) using D’Alembert’s principle and determine the natural frequencies, periods,
normalized modal matrix, and participation factors using MATLAB.
Multi-Degree-of-Freedom System Analysis 201
50 psf
u2(t)
W10×21
12 ft
I = 106.3 in4
100 psf
u1(t)
W10×45 15 ft
I = 248.6 in4
30 ft
7.7 Given the following mode shapes and frequencies, compute the participation factors for a
3-story building with floor weights of 120 kips for the first floor, and 80 kips for the second
floor and roof.
k1 = 40 kips/in 15 ft
7.9 Given a five-story building frame subjected to a ground acceleration due to a 0.25g PGA
earthquake characterized by the 84.1% design spectrum, use an MDOF analysis to deter-
mine (i) peak displacements, (ii) maximum base shear, and (iii) maximum overturning
moments. k = 326.3 kips/in, w5 = 50 kips, and the weights on the other four floors equal
100 kips. Assume rigid beams and damping of 5%.
REFERENCES
Chopra, A. K., Dynamics of Structures: Theory and Applications to Earthquake Engineering, 4th edition,
Prentice-Hall, Upper Saddle River, NJ, 2012.
Villaverde, R., Fundamental Concepts of Earthquake Engineering, CRC Press, Boca Raton, FL, 2009.
8 Seismic Code Provisions
Design of structures must be based on a minimum standard of care to safeguard the safety,
health, and welfare of the public. Municipal, state, or federal governments concerned with the
safety of the public have established building codes used to control the construction of struc-
tures within their jurisdiction. These design codes specify among other things the design loads.
While building codes may vary from city to city, most municipalities rely on regional codes and
design standards for the specification of loads. These codes and standards include general build-
ing codes, such as the International Building Code (IBC) and the ASCE/SEI (2010) “Minimum
Design Loads for Buildings and Other Structures” standard, from here on referred to as ASCE-7.
These codes and standards are developed by various organizations and present the best opinion
of those organizations as to what constitutes good practice. In this chapter, we provide an over-
view of structural design philosophies and explain how seismic loads are incorporated as part
of structural design. The primary focus of this chapter is to describe the parameters associated
with determining earthquake loads in accordance with ASCE-7 and to demonstrate the proce-
dures for determining earthquake loads using the ELFP, seismic response history analysis, and
MRS analysis.
1. A structural failure (or strength) limit state is the more critical of the two because it is
concerned with safety and it relates to the maximum load carrying capacity of a structure.
2. A serviceability limit state relates to structural performance under normal service
conditions.
Both design philosophies focus on the ultimate limit state and allow the designer some freedom
of judgment in serviceability. The main concern from a strength limit state standpoint is ensuring
that the capacity (resistance, R) of the structure exceeds the demand (effect of loads, Q). Since it is
203
204 Introduction to Earthquake Engineering
economically impracticable to design 100% safe structures, we must assess the risk of failure, or
the reliability of the structure. To assess the reliability of a structure, we must understand how the
uncertainties on material properties, geometry, and material durability relate to the load uncertain-
ties; not all the extremes of loads can occur simultaneously on the structure, or, at least, it is very
unlikely. In ASD, all uncertainties are accounted using a single quantity, the factor of safety (FS),
while the LRFD uses a rational approach to obtain a more uniform assessment of the reliability of a
structure. Thus, the main difference between ASD and LRFD is that LRFD rationally accounts for
the statistical variability in the loads (and materials), and is based on reliability analysis to obtain a
probability-based assessment of structural safety.
R (8.1)
FS ≥
∑Q
It is important to note that both sides of the inequality must be evaluated for the same conditions.
For example, load effect that produces tensile stresses should be compared with the tensile strength
of the member in question. The strength of a member is material dependent; a properly designed
system of members must ensure vertical and lateral load-resisting system integrity by providing
ductility, energy dissipation, and avoiding progressive collapse by providing redundancy and pre-
serving the load path—all while maintaining deformations within the prescribed limits.
The stress produced by the loads, ΣQ, is determined once the structural form has been selected
using combinations of loads that can be reasonably expected to act on the structure. That is, not
all possible loads will act concurrently; thus, ASCE-7 (Section 2.4) specifies the following load
combinations:
1. D
2. D + L
3. D + (Lr or S or R)
4. D + 0.75 L + 0.75 (Lr or S or R)
5. D + (0.6 W or 0.7 E)
6a. D + 0.75 L + 0.75(0.6 W) + 0.75 (Lr or S or R)
6b. D + 0.75 L + 0.75(0.7 E) + 0.75 S
7. 0.6 D + 0.6 W
8. 0.6 D + 0.7 E
where:
D represents the dead loads, which are loads that have constant magnitude, remain in one posi-
tion, and are usually known to a high degree of certainty. They include the weight of vari-
ous structural elements and weight of objects permanently attached to the structure, such
as walls, roofs, ceilings, and equipment
L represents the live loads, which are those loads that vary in magnitude and position with time.
They are caused by the structure being occupied, used, and maintained. In general, live
loads are induced by gravity, and can be stationary or transient, which may still be treated
as a static load, unless applied rapidly, in which case, the dynamic effect creates additional
inertial load called impact
Seismic Code Provisions 205
Lr represents the roof live loads, which are loads produced by repair/maintenance workers, their
equipment and materials, or during the life of the structure by movable objects such as
planters. There is a limit to the amount of load that can be realistically placed upon roofs
because most roofs are sloped
S represents the snow load, which depends on ground snow weight, the building’s general shape, and
roof geometry (flat roofs are subjected to higher loads compared to sloped roofs)
R represents the rain load, which depends on the functionality of the drainage facilities, and
only a concern when rainwater accumulates faster than it runs off, ponding. According
to ASCE-7, elements should be designed to support all rainwater that accumulates when
primary drains are blocked and the water rises above the inlet of the secondary drainage
system
W represents the wind load, which is a lateral load caused by wind pressure resulting from the
kinetic energy of moving air as it strikes an object in its path
E represents the earthquake load, which is the main subject of this book
These loads are actions that result from the weight of all building materials, occupants and their
possessions, environmental effects, and are categorized based on their character and duration.
The history of each of these loads can be represented schematically as shown in Figure 8.1.
Dead load
Time
Sustained
live load
Several years
Time
Transient
Several
live load
hours/days
Time
Seconds
Earthquake
n years
Time
Wind
3s
Time
Load combos
Time
D D+L D + 0.75 L + 0.75(0.6 W) D + 0.75 L + 0.75(0.7 E)
FIGURE 8.1 Schematic history of loads and their superposition to obtain the load combinations.
206 Introduction to Earthquake Engineering
This figure shows that the dead load remains constant over its entire history, the live load can be
divided into a sustained load (one that may have variations at different points in its history, such
as building partitions) and a transient load (one that changes frequently, both in duration and
magnitude), the earthquake load is based on short events as was discussed in earlier chapters, and
a wind load that includes sustained intensity and wind gusts. As shown in the figure, the largest
magnitudes of each of these loads are not likely to occur simultaneously, which is why various
load combinations are needed to represent potential critical loadings at different points during the
life of a structure.
Gravity-induced loads on any floor system are assumed to act as a uniform pressure on the floor
slab. We are only concerned with the structural system; so to analyze the effects of these pressure-
like forces on each structural member, we need to rationally distribute them to the supporting mem-
bers. To accomplish this, we use the concept of the tributary area, the area of the slab that is carried
by a particular structural member.
φR ≥ ∑γ Q
i i (8.2)
Again, both sides of the inequality must be evaluated for the same conditions. The left-hand side
of this inequality is the available strength, while the right-hand side is the required strength, which
is computed using the following load combinations (ASCE-7, Section 2.3):
1. 1.4 D
2. 1.2 D + 1.6 L + 0.5 (Lr or S or R)
3. 1.2 D + 1.6 (Lr or S or R) + (L or 0.5 W)
4. 1.2 D + W + L + 0.5 (Lr or S or R)
5. 1.2 D + E + L + 0.2 S
6. 0.9 D + W
7. 0.9 D + E
where all quantities were previously defined. Also, in this design philosophy, unfactored loads are
used for serviceability requirements because safety is not usually of concern for serviceability.
d. Dual systems with special moment frames (in steel and concrete)
e. Dual systems with intermediate moment frames (in steel and concrete)
f. Ordinary shearwall frame interactive systems (in concrete)
g. Cantilever column systems (in steel, concrete, and timber)
h. Steel systems not detained for seismic resistance
Some of these systems are not permitted in seismic-resistant design; and those permitted are sub-
ject to building height and other limitations. These limitations depend on the various seismic
design categories (SDCs, discussed later in Section 8.4.2, and ASCE-7, Section 11.6), which are
classifications of building requirements based on a desired performance that depends on a build-
ing’s occupancy, the effects of probable ground-shaking intensity at the building site, and struc-
tural irregularities.
In all cases, damage depends on: intensity and duration of ground shaking, building configuration
(irregular buildings are more susceptible to damage), lateral force-resisting system type (e.g., shear-
walls, braced or unbraced frames), building material (e.g., concrete, steel, wood, or masonry), and
quality of construction. Note that IBC or ASCE-7 provides no guidance for protection against earth
movement, earth slides, liquefaction, or direct fault displacement. The ASCE-7 seismic loading-
related specifications are provided in 13 chapters, the titles of which are as follows:
This chapter only covers some of the seismic code provisions, mainly procedures described in
Chapters 11, 12, and 16 of ASCE-7.
208 Introduction to Earthquake Engineering
where:
Ev is not a peak value; rather it recognizes that the peak values of the horizontal and vertical
seismic forces are unlikely to occur simultaneously. Also, Ev need not be included when
SDS ≤ 0.125g
ρ is the redundancy factor and is discussed in Section 8.3.1
QE is the effect of horizontal seismic forces (exhibited as internal forces in the axial, shear, and
flexure members), and will be discussed in Sections 8.6−8.8
D is the dead load previously defined
SDS is the design spectral acceleration for short periods, and will be discussed in Section 8.5
Ωo is the overstrength factor discussed in Section 8.3.2
Regarding the direction of the design seismic loads, ASCE-7 (Section 12.5) specifies that E be
applied in directions that produce the largest effect on structural members. The code does, however,
provide guidelines to satisfy this requirement. For example, design seismic forces are permitted to
be applied separately and independently in each of two orthogonal directions for structures assigned
to SDC B (discussed in Section 8.4.2) and regular structures in all other categories.
EXAMPLE 8.1
A column of a steel special concentrically braced frame in a single story medical office build-
ing (SDC = D) supports an axial dead load, D = 35 kips, and horizontal seismic load effect,
QE = ±15 kips. Given SDS = 1.25, an overstrength factor, Ωo = 2.5, and a redundancy factor,
Seismic Code Provisions 209
ρ = 1.3, determine the maximum and minimum axial forces in the column using LRFD load com-
binations. Repeat the problem accounting for overstrength.
Solution
1. Determine the maximum axial load. Determine the maximum earthquake load, E, using
Equation 8.3:
Thus, taking the positive quantity in the first term, the maximum earthquake load is
. D + E + L + 0.2S = 12
5) 12 . (35 kips) + 28.3 kips = 70.3 kips
2. Determine the minimum axial load. Determine the minimum earthquake load, E, using
Equation 8.4:
E = Eh – Ev = ρQE – 0.2SDSD = 13
. (±15 kips) – 0.2(125
. )(35 kips) = ±19.5 kip
ps – 8.8 kips
Thus, taking the negative quantity in the first term, the minimum earthquake load is
3. Determine the maximum axial load including overstrength. Determine the maximum
earthquake load, Em, using Equation 8.5:
Thus, taking the positive quantity in the first term, the maximum earthquake load is
. D + Em + L + 0.2S = 12
5) 12 . (35 kips) + 46.3 kips = 88.3 kips
4. Determine the minimum axial load including overstrength. Determine the minimum
earthquake load, Em, using Equation 8.6:
Em = Eh – Ev = ΩoQE – 0.2SDSD = 13
. (±15 kips) – 0.2(125
. )(35 kips) = ±19.5 kips – 8.8 kips
210 Introduction to Earthquake Engineering
Thus, taking the negative quantity in the first term, the minimum earthquake load is
8.3.1 Redundancy Factor, ρ
Redundancy is a characteristic in which multiple paths of resisting lateral loads are provided. ρ is
intended to encourage system redundancy; when redundancy is lacking, ρ in Equations 8.3 and 8.4
can result in a penalty, or increase in the seismic load of up to 30%. Also, ρ is assigned to the seis-
mic force-resisting system in each of two orthogonal directions for all structures. For the following
conditions, ρ can be taken as 1.0, and thus there is no penalty:
The redundancy factor should be taken as 1.3 (a 30% penalty) for structures assigned to SDC D, E,
and F. Unless one of the following two conditions is met, whereby ρ is permitted to be taken as 1.0:
1. Stories resisting forces larger than 35% of the base shear in any given direction that com-
plies with Table 12.3-3 in ASCE-7.
2. Structures that are regular in plan at all levels and are provided with at least two bays of
seismic force-resisting perimeter framing on each side of the structure, in each orthogonal
direction, with each story resisting more than 35% of the base shear. For masonry, concrete,
or steel plate shearwalls, the number of bays shall be calculated as the length of shearwall
divided by the story height. For light-framed construction shearwalls, the number of bays
shall be calculated as two times the length of shearwall divided by the story height.
EXAMPLE 8.2
A single-story building shown in Figure E8.1 with an average roof height of 12 ft is constructed
with a flexible roof diaphragm assigned to SDC D. The base shear was determined to be 100 kips,
1 2
100 ft
SFRS
50 ft
which can be resolved into total forces along wall lines 1 and 2 of 50 kips each. Determine the
redundancy factor for each of three lateral force-resisting systems consisting of (1) 18-foot long
concrete shearwalls, (2) single-steel ordinary concentrically braced frames, and (3) double-steel
special moment frames.
Solution
1. 18-foot long concrete shearwalls.
Determine the number of bays using one of the following relationships: Light-frame con-
struction number of bays is given as
where:
Lw is the length of the shearwall
hsx is the story height
#bays = Lw /hsx
Therefore, ρ = 1.3
3. Double-steel special moment frames.
Therefore, ρ = 1.0
Base shear, V
Elastic response spectrum
VE
VY = ΩoVS
VS = VE I ∆D
Cd =
R ∆S
∆S ∆D ∆E Drift, ∆
Period, T
EXAMPLE 8.3
Use ASCE-7, Table 12.2-1, to determine seismic force-resisting system parameters and limitations
for a steel special moment frame (SSMF) depicted in Figure E8.2. Also, list an advantage and a
disadvantage for SSMF systems.
Seismic Code Provisions 213
Solution
1. Determine seismic force-resisting system parameters. From ASCE-7, Table 12.2-1, the
SSMF system is part of group C, C.1. The values of design coefficients and factors are as
follows:
Response modification coefficient, R = 8
The overstrength factor Ωo = 3
The deflection amplification factor Cd = 5.5
2. Identify the limitations for the seismic force-resisting system. From ASCE-7, Table 12.2-1,
group C, C.1, structural system limitations are available according to SDC as
SDC B C D E F
Restriction NL NL NL NL NL
For each SDC of the SSMF, NL = not limited is listed, which implies that the system is
permitted everywhere and has no height limits.
3. Describe an advantage and a disadvantage for SSMF systems.
Advantage: because of the high R value, the base shear is relatively low.
Disadvantage: because of the high Cd value, the drifts are relatively high.
Stockton, CA
2% Probability of being
exceeded in 50 years
FIGURE 8.3 Maximum considered ground motion intensity in San Joaquin County, California (as percent
of acceleration due to gravity).
SS and S1 are presented in map form in ASCE-7, Figures 22-1–22-6, assuming 5% damping and a
site underlaid by rock, see Figure 8.3, for the contour maps in San Joaquin County, California. S1
represents acceleration for a period of 1.0 s, while SS represents acceleration for a period of 0.2 s
(short period). These maps are difficult to read, especially in cases where contour lines are closely
spaced. A more accurate and convenient approach is to use the USGS Seismic Hazards Mapping
Application, available at: http://earthquake.usgs.gov/designmaps/us/application.php, as shown in
Figure 8.4. This application generates detailed reports that provide a site-specific response spec-
trum; see Example 8.4.
As indicated in Chapter 5, Figure 5.13, the response spectrum is affected by local soil condi-
tions. Since SS and S1 are based on rock subsurface conditions, they must be adjusted for other
FIGURE 8.4 USGS seismic hazard mapping application input screen. (Courtesy of USGS.)
Seismic Code Provisions 215
types of soil conditions. ASCE-7, Chapter 20, classifies a site as Site Class A (hard rock), B (rock),
C (very dense soil and soft rock), D (stiff soil), E (soft soil), or F (soil) based on soil shear wave
velocity, vs; if vs is not known, ASCE-7 allows the use of standard penetration resistance or und-
rained shear strength values. This information is typically included as part of a geotechnical
engineering report. If no soil properties are known, ASCE-7 allows the use of Site Class D as a
conservative assumption.
The response spectrum associated with MCE SS and S1 can be adjusted to include site class
effects, SMS and SM1; the results are given as
S MS = Fa SS and S M 1 = Fv S1 (8.7)
where:
Fa is a site coefficient per ASCE-7, Table 11.4-1, that depends on soil type and ranges in values
from 0.8 for rock to 2.5 for soft soil
Fv is a site coefficient per ASCE-7, Table 11.4-2, that also depends on soil type and ranges in
values from 0.8 for rock to 3.5 for soft soil
SMS is the mapped risk-targeted maximum considered earthquake (MCER) spectral response
acceleration parameter at short periods
SM1 is the mapped MCER spectral response acceleration parameter at a period of 1 s
Also, the values Fa and Fv are smaller for large spectral accelerations, which is intended to cap-
ture the nonlinear behavior of soils that prevent ground motion amplifications.
The response spectrum associated with SMS and SM1 (based on MCER ground motion) is intended
to produce the most severe earthquake effects. MCE ground motion corresponds to a 2% probabil-
ity of being exceeded in a 50-year period (∼2%/50), which corresponds to an approximate return
period, or a recurrence interval of 2500 years, as obtained in Section 2.4. However, the nominal
design level is based on a maximum design earthquake (MDE) ground motion, which corresponds
to a 10% probability of being exceeded in a 50-year period, or approximately a 500-year recurrence
interval (∼10%/50). ASCE-7 accounts for two-thirds of the MCE ground motion spectrum accelera-
tions, SMS and SM1, in order to reduce the MCE ground motion from 2%/50 to 10%/50 hazard level;
the results are given as
S DS = 2 / 3 S MS and S D1 = 2 / 3 S M 1 (8.8)
where:
SDS is the design, 5% damped, spectral response acceleration parameter at short periods
SD1 is the design, 5% damped, spectral response acceleration parameter at a period of 1 s
The design response spectrum associated with SDS and SD1 can be obtained using ASCE-7,
Section 11.4.5. Draw the design spectral response acceleration, Sa, as a piecewise continuous func-
tion, with four segments as shown in Figure 8.5 or algebraically as follows:
For
T
T ≤ To , Sa = S DS 0.4 + 0.6
To
To = 0.2 Ts
216 Introduction to Earthquake Engineering
SDS
SD1
Sa =
SD1 T SD1TL
Sa =
T2
SD1
Ts =
SDS
To = 0.2Ts Ts 1.0 TL
Structure period, T
Ts = SD1/SDS, which represents the transition period from acceleration- to velocity-controlled seg-
ments of the design spectrum (see Figure 5.8).
T is the fundamental period of the structure.
Note that for T = 0, the spectral acceleration, Sa, is approximately equal to 0.4SDS, which repre-
sents the design-level peak ground acceleration.
EXAMPLE 8.4
Determine the spectral response acceleration parameters and draw the response spectrum for a
site located on the campus of University of the Pacific in Stockton, California, which will be used
for the construction of an office building. The site (soil) is class D.
Solution
1. Determine the maximum considered spectral accelerations for short and 1 s periods, SS
and S1. From the seismic intensity maps in ASCE-7, Chapter 22, SS and S1 (as fractions of
acceleration due to gravity) are estimated from Figure 8.3,
SS = 0.85
S1 = 0.30
2. Determine the site coefficients for the given site class. For Site Class D, using ASCE-7,
Tables 11.4-1 and 11.4-2, identify the site coefficients.
Fa = 1.16 using linear interpolation of values in Tables 11.4-1
Fv = 1.80 using Tables 11.4-2
3. Determine the maximum considered spectral accelerations, adjusted for site effects, for
short and 1 s periods, SMS and SM1,
Seismic Code Provisions 217
SMS = Fa SS = 1.16(0.85) = 0.896
SM1 = Fv S1 = 1.80(0.30) = 0.54
4. Determine the design spectral accelerations for short and 1 s periods, SDS and SD1,
SDS = 2/3SMS = 2/3(0.896) = 0.65
SD1 = 2/3SM1 = 2/3(0.54) = 0.36
5. Draw the design response spectrum. The design response spectrum, Sa, is a piecewise
continuous function with four segments:
T T
For T ≤ To , Sa = SDS 0.4 + 0.6 = 0.650.4 + 0.6 = 0.26 + 3.5 T
To 0.11 s
where the transition periods To and Ts are:
To = 0.2Ts = 0.11
SD1 0.36
Ts = = = 0.55 s
SDS 0.65
The long-period transition period TL obtained from ASCE-7, Figure 22-12, is equal to 8 s. Also,
notice that for T = 0, the spectral acceleration is approximately equal to 0.26, which represents the
design-level peak ground acceleration. Figure E8.3 shows the entire spectrum.
The sample report for this example from the USGS Seismic Hazards Mapping Application at http://
earthquake.usgs.gov/designmaps/us/application.php also provides the response spectrum (Figure E8.4).
0.6
0.5
0.4
Sa (g)
0.3
0.2
0.1
0
0 1 2 3 4 5 6 7 8 9 10
T (s)
FIGURE E8.3 Acceleration response spectrum at the University of the Pacific, Site Class D.
218 Introduction to Earthquake Engineering
To determine the SDC for a structure using Table 8.1 (the most severe case of SDS, SD1, or S1 governs),
we must determine the soil site class using ASCE-7, Chapter 20, the MCE spectral response accel-
erations (SS and S1), the design spectral response accelerations (SDS and SD1) as described in Section
8.4.1, and finally the occupancy category using ASCE-7, Table 1.5-1 (discussed in the following
paragraphs), or use the USGS application with results provided in the detailed report.
ASCE-7 also classifies buildings based on the occupancy category and importance factor for the
building system. Occupancy categories are assigned to various buildings and other structures based
on their use and nature of occupancy (see ASCE-7, Table 1.5-1), and they are intended to increase
the safety level of essential facilities and those that pose a significant risk to human life. The nomi-
nal design level is Occupancy Category II, which corresponds to an MDE ground motion (10%/50).
The higher risk levels correspond to the probabilities listed in Table 8.2. As this table shows, essen-
tial and critical structures are designed for an MCE ground motion (2%/50), which is the level given
by the mapped spectrum parameters in ASCE-7, Chapter 22. Thus, the importance factors listed in
Table 8.2 effectively adjust the design response spectrum for various hazard levels.
The importance factors indirectly address the objectives of IBC seismic provisions discussed in
Section 8.2, which are related to the performance of systems for various levels of earthquake shak-
ing, that is,
TABLE 8.1
Seismic Design Categories; Summary of ASCE-7, Section 11.6
Based on SDS (Table 11.6-1) Based on SD1 (Table 11.6-2) I or II or III IV
SDS < 0.167 SD1 < 0.067 A A
0.167 ≤ SDS < 0.33 0.067 ≤ SD1 < 0.133 B C
0.33 ≤ SDS < 0.50 0.133 ≤ SD1 < 0.20 C D
SDS ≥ 0.50 SD1 ≥ 0.20 D D
S1 ≥ 0.75 E F
220 Introduction to Earthquake Engineering
TABLE 8.2
Occupancy Categories and Importance Factors
Occupancy
Category Nature of Occupancy Return Period (Years) Importance Factor, Ie
I Low hazard structures 475 1.00
(10% probability of exceedance in 50 years)
II Standard occupancy, other than I, III, or IV 475 1.00
(10% probability of exceedance in 50 years)
III Assembly structures 975 1.25
(5% probability of exceedance in 50 years)
IV Essential or critical structures 2475 1.50
(2% probability of exceedance in 50 years)
Occupancy Categories I and II provide the minimum level of protection and are intended to
address the risk of structural failure and loss of life, but not to limit structural damage, whereas
Occupancy Categories III and IV provide a higher level of protection against loss of life and prop-
erty by limiting the damage to a minimum in order to provide a continued function of facilities
during and after a seismic event.
The analysis procedures allowed by the code include ASCE-7, Section 12.8, ELFP as well as two
dynamic methods: ASCE-7, Section 12.9, MRS analysis, and ASCE-7, Chapter 16, seismic response
history procedures, all of which will be discussed in this chapter. The dynamic methods are permit-
ted for all building systems in all six SDCs. The ELFP is permitted for all building systems in SDCs
A, B, and C, and is allowed for most cases in SDCs D, E, and F, except buildings exceeding 160 ft
in height with T > 3.5 Ts and those with T < 3.5 Ts and horizontal irregularities 1a or 1b, or vertical
irregularities 1a, 1b, 2, or 3 (as noted in ASCE-7, Table 12.6-1). Even for the exempt cases, the ELFP
is used in scaling the results of the dynamic methods.
Building irregularities are discussed in ASCE-7, Section 12.3.2, and are defined as horizontal
(ASCE-7, Table 12.3-1) or vertical (ASCE-7, Table 12.3-2). Horizontal irregularities include: (1a)
torsional, (1b) extreme torsional, (2) reentrant corners, (3) diaphragm discontinuities, (4) out-of-
plane offsets, and (5) nonparallel resisting systems. Vertical irregularities include: (1a) stiffness-soft
story, (1b) stiffness-extreme soft story, (2) weight (mass), (3) vertical geometric, (4) In-plane discon-
tinuity in vertical lateral force-resisting element, (5a) discontinuity in lateral strength–weak story,
and (5b) discontinuity in lateral strength−extreme weak story.
1. Compute seismic weight W using details presented in Section 3.1.2 and repeated here for
convenience (ASCE-7, Section 12.7.2).
Seismic Code Provisions 221
where:
DL is the dead load of the structural system that is tributary to each floor
StL is the storage load, which is a live load in areas used for storage, such as a
warehouse
PL is the partition load, when applicable
WPE is the operating weight of permanent equipment
SL is the flat roof snow load when it exceeds 30-psf, regardless of roof slope
LaL is the landscape loads associated with roof and balcony gardens
2. Compute the natural period of the structure, T, following ASCE-7, Section 12.8.2. T is the
calculated fundamental (first mode) period of a structure in the direction under consider-
ation. It can be determined using the structural characteristics of the lateral force-resisting
elements as introduced in Section 3.1.1, or the generalized SDOF properties covered in
Chapter 6, or a full eigenanalysis covered in Chapter 7. When one of these methods is used
to determine T, it is subject to lower and upper limits as follows:
Use T if Ta ≤ T ≤ CuTa
T = Ta if T < Ta
where:
Cu is the coefficient for the upper limit on the calculated period from ASCE-7,
Table 12.8-1
Ta is the approximate fundamental period (ASCE-7, Section 12.8.2.1)
Alternatively, the code permits the use of Ta, calculated using the details presented in
Section 3.1.4, and repeated here for convenience,
Ta = Ct hnx (8.9)
where:
hn is the height in feet above the base to the highest level of the structure
Ct and x are determined from Table 3-2 (or ASCE-7, Table 12.8-2)
Ta is based on the measured response of buildings in high seismic regions; its value can
be adjusted to get a more accurate approximation of the actual period by including local
seismicity using Cu to obtain the upper bound period, CuTa. This is based on the best-fit of
the measured response.
3. Computer seismic base shear (ASCE-7, Section 12.8.1). The base shear is caused by the
inertial force (mass times acceleration) from the earthquake motion, F = ma = W(a/g),
where g is the acceleration due to gravity. ASCE-7 modifies this expression to obtain an
equivalent static lateral force acting at the base of the structure, the base shear, V:
V = C sW (8.10)
222 Introduction to Earthquake Engineering
where:
W is the seismic weight computed in step 1
Cs is the seismic response coefficient, expressed as a fraction of g (ASCE-7, Section
12.8.1.1)
S DS
Cs = (8.11)
R /I e
where:
SDS is the design spectral response acceleration in the short period range, computed
using the first of Equation 8.8 (ASCE-7, Section 11.4.4)
R is the response modification factor in ASCE-7, Table 12.2-1
Ie is the importance factor in Table 8.2 (ASCE-7, Section 11.5.1)
The value of Cs need not exceed
S D1 S D1TL
Cs = for T ≤ TL or Cs = for T > TL (8.12)
T ( R /I e ) T 2 ( R /I e )
Collectively, these three relationships for Cs represent the inelastic design response spec-
trum (Equation 8.11 is the constant acceleration portion and Equations 8.12 represents the
constant velocity and displacement portions, respectively), obtained by dividing the elastic
design response spectrum (shown in Figure 8.5) by the factor R/Ie. Figure 8.6 provides an
illustration of the resulting inelastic design response spectrum.
Also,
or
0.5S1
Cs > , for S1 ≥ 0.6 g (8.14)
R /I e
SDS
(R/Ie) SD1 Cs (min) = 0.044SDSIe
T(R/Ie) or 0.01
0.5S1
or , for S1 ≥ 0.6g
R/Ie
Cs(g)
SD1 SD1TL
Ts =
SDS
T 2(R/Ie)
Ts TL
T (s)
where:
SD1 is the design spectral response acceleration at a period of 1 s, computed using the
second of Equations 8.8 (ASCE-7, Section 11.4.4)
T is the fundamental period of the structure computed in step 2
TL is the long-period transition period (ASCE-7, Section 11.4.5)
S1 is the MCE spectral response acceleration at a period of 1 s discussed in Section
8.4.1 (ASCE-7, Section 11.4.4)
The two limits in Equation 8.13 are intended to represent the minimum base shear force
levels to safeguard against the collapse of long-period structures (uncertainty related to
P-Δ effects), while Equation 8.14 accounts for the effects of near-fault directivity.
4. Compute the equivalent lateral forces (ASCE-7, Section 12.8.3). The lateral seismic force,
Fx, induced at any level x as illustrated in Figure 8.7, can be determined using the following
relationship:
Fx = CvxV (8.15)
where:
V is the base shear computed in step 3
Cvx is the vertical distribution factor
wx ⋅ hxk
Cvx = (8.16)
∑
n
wi ⋅ hik
i=1
where:
wx and wi are the weights at level x or i, depicted in Figure 8.7
h x and hi are the heights from the base to the level x or i, depicted in Figure 8.7
n is the number of stories
k depends on the structural period and accounts for higher mode effects
k = 1 for T ≤ 0.5 s
F4 w4 h4
k=2
F3 w3 h3
F2 w2 h2
k=1
F1 w1 h1
V
Mb
These formulas are approximately the same as Equation 6.37 for a generalized SDOF
shear building, where we assumed that the lateral response is dominated by the funda-
mental mode. The shape vector, ψi, in this case is assumed to vary from linear (k = 1) to
parabolic (k = 2), as shown in Figure 8.7. The parabolic mode is intended to account for
the response of flexible structures.
5. Compute overturning moment (ASCE-7, Section 12.8.5). As discussed in Section 6.1.3,
with the equivalent static forces, we can determine the overturning moment by applying
static equilibrium to the free-body diagram depicted in Figure 8.7, that is,
Mb = ∑h F i i (8.17)
i=1
As shown in Figure 6.3, we can also obtain the internal story shear force, Vmax i, and the
internal story moment, Mmax i, at an arbitrary level i by applying static equilibrium to the
right-hand side free-body diagram as shown by Equations 6.38 and 6.39, respectively, and
are repeated as follows for convenience:
Vmax i = ∑f
j =i
jo
M max i = ∑ (h − h ) f
j =i
j i jo
Again, in the derivation of Equations 8.17 and 6.39, we assumed that the floor weights are
directly in the center of the building frame, thus making no contribution to the moment
equilibrium equation. However, for buildings with off-center weights, a full moment static
equilibrium analysis is conducted to account for the contribution of these off-center forces.
6. Determine story drifts (ASCE-7, Section 12.8.6). With the lateral forces, we can perform
a structural analysis to determine lateral displacements. The relative displacement of
each story at the center of mass is defined as the drift. For cases where centers of mass
do not align and for those cases with irregularities, restrictions are specified in ASCE-7.
Note that the loads used to calculate drifts are at strength level (not service), as indicated
in ASCE-7, Figure 12.8-2. The drift is computed using the following inelastic displacement
at level x:
Cdδxe
δx = (8.18)
Ie
where:
Cd is the deflection amplification factor discussed in Section 8.3.2 and given in ASCE-7,
Table 12.2-1
Seismic Code Provisions 225
∆ x = δx − δx−1 (8.19)
where:
δ x is the deflection at the top of level x determined using Equation 8.18
δ x−1 is the deflection at the bottom of level x determined using Equation 8.18
To ensure the stability of the overall system (i.e., avoid P-Δ effects) and to minimize dam-
age to nonstructural elements, these drifts must be kept within allowable limits per ASCE-7,
Section 12.12.1, which are listed in Table 12.12-1 and depend on the structure type and
occupancy category.
When P-Δ effects are significant, provisions in ASCE-7, Section 12.9.6, should be followed to
ensure stability. For cases where torsional effects are significant, ASCE-7, Section 12.8.4.3, requires
the application of torsional amplification factors in the analysis.
EXAMPLE 8.5
Given the following three-story office building frame located on the campus of University of
the Pacific in Stockton, California, compute the approximate structural period Ta, the design
base shear V, the design earthquake loads acting on each floor Fx, and the overturning moment
at the base. The building in plan view is square; the structural system being proposed is a
special steel concentrically braced frame (SSCBF) as shown in Figure E8.5 with an estimated
weight on each of the first two floors of 100 and 80 kips on the roof. The geotechnical engineer
has classified the site (soil) class as D.
Solution
1. Determine the seismic weight. The weight of the roof and two floors is given; the total
weight of the system is the sum of these weights,
2. Determine the period. No information about structural stiffness is given, so we use the
approximate period given by Equation 8.9,
Ta = Ct hnx
12 ft
12 ft
12 ft
FIGURE E8.5 Elevation view of the special steel concentrically braced frame.
226 Introduction to Earthquake Engineering
where:
hn = 36 ft and is the building height
for an SSCBF, which corresponds to all other structural systems in Table 3.2,
Ct = 0.02
x = 0.75
Thus,
Note that this value is very close to the result of Equation 3.19, Ta = 0.1N = 0.1(3) = 0.3 s.
This relationship can be used to check the validity of the calculated values of the period.
3. Calculate the base shear, V = CsW per ASCE-7, Section 12.8. The base shear is obtained
using Equation 8.10, with Equations 8.11 and 8.12, or the response spectrum shown in
Figure 8.6. Using the spectral values obtained in Example 8.4, draw the response spec-
trum for this case as shown in Figure E8.6.
With Ta = 0.29 s, we can enter this spectrum. Since To < Ta ≤ Ts, Sa = SDS = 0.65 and
SDS
Cs =
R /Ie
As noted in the problem statement, the building will be used for office space, which
corresponds to an occupancy category of II; thus, the importance factor obtained from
Table 8.2 is
Ie = 1.0
Also, we can determine the SDC using Table 8.1 (or the USGS application discussed in
Example 8.4) to ensure that the selected seismic load-resisting system is permitted, that
0.6
0.5
0.4
Sa (g)
0.3
0.2
0.1
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Ta T (s)
is, for Occupancy Category II, and SDS = 0.65 > 0.5 and SD1 = 0.36 > 0.2, the governing
SDC is D, as shown in the following table.
SDS < 0.167 SD1 < 0.067 A A
0.167 ≤ SDS < 0.33 0.067 ≤ SD1 < 0.133 B C
0.33 ≤ SDS < 0.50 0.133 ≤ SD1 < 0.20 C D
SDS ≥ 0.50 SD1 ≥ 0.20 D D
S1 ≥ 0.75 E F
Thus, for SSCBF (ASCE-7, Table 12.2-1, case 2 in part B), the height limit is 160 ft, which
is larger than the proposed 36 ft building height. The seismic response coefficient is also
obtained from ASCE-7, Table 12.2-1,
R = 6.0
SDS 0.65
V = C sW = W= ( 280 kips) = 30.7 kips
R /Ie 6.0 /1.0
4. Determine lateral forces at each level of the structure, Fx = CvxV per ASCE-7, Section
12.8.3. The forces are obtained using Equations 8.15 and 8.16 for each level, x:
Fx = CvxV
where:
V = 30.7 kips, computed in step 3
Cvx is the vertical distribution factor with k = 1 since Ta = 0.29 s ≤ 0.5 s
w x ⋅ hx w x ⋅ hx w x ⋅ hx
Cvx = = =
∑
3
w i ⋅ hi 100 kips(12 ft ) + 100 kips( 24 ft ) + 80 kips(36 ft ) 6480 k ⋅ ft
i =1
w1 ⋅ h1 100 kips(12 ft )
F1 = Cv1V = (30.7 kips) = (30.7 kips) = 5.68 kips
6480 k ⋅ ft 6480 k ⋅ ft
w 2 ⋅ h2 100 kips( 24 ft )
F2 = Cv 2V = (30.7 kips) = (30.7 kips) = 11.36 kips
6480 k ⋅ ft 6480 k ⋅ ft
w 3 ⋅ h3 80 kips(36 ft )
F3 = Cv 3V = (30.7 kips) = (30.7 kips) = 13.63 kips
6480 k ⋅ ft 6480 k ⋅ ft
5. Determine the overturning moment using static equilibrium. With the lateral forces com-
puted in the last step, we can determine the overturning moment at the base using
Equation 8.17,
3
Mb = ∑ h F = 5.68 kips(12 ft) + 11.36 kips(24 ft) + 13.63 kips(36 ft)) = 831kip ⋅ ft
i i
i =1
228 Introduction to Earthquake Engineering
13.63 kips
12 ft
11.36 kips
13.6 –490
12 ft
5.68 kips
25 –763
12 ft
FIGURE E8.7 Summary of analysis results using the equivalent lateral force method.
We can also obtain the internal story shear forces and moments at each level using
Equations 6.38 and 6.39, respectively. Alternatively, we can draw shear force and bend-
ing moment diagrams by treating the building as a cantilever beam in order to obtain
the internal story shear forces and moments. Figure E8.7 summarizes the results of the
analysis.
EXAMPLE 8.6
Check the allowable story drifts for the three-story office building frame given in Example 8.5.
A structural analysis of the building frame was conducted with lateral forces determined in Example
8.5 in order to determine the elastic displacements at each level (δ1e = 0.65 in, δ2e = 1.3 in, and
δ3e = 1.6 in) and illustrated in Figure E8.8.
SOLUTION
1. Determine inelastic displacements per ASCE-7, Section 12.8.6. Compute inelastic dis-
placements at each level x using Equation 8.18,
Cd δxe
δx =
Ie
where:
Cd = 5.0 for steel special concentrically braced frames given in ASCE-7, Table 12.2-1
Ie = 1 for the office building
δxe is the deflection at level x given in Figure E8.8.
FIGURE E8.8 Lateral forces and elastic displacements at each level of the building frame.
Seismic Code Provisions 229
2. Determine story drifts. Compute drifts at each level using Equation 8.19,
3. Check the allowable story drifts per ASCE-7, Section 12.12.1. For buildings in Occupancy
Category II, ASCE-7, Table 12.12-1, gives an allowable drift (for four stories or less struc-
tures with interior walls, partitions, etc. designed to accommodate story drifts) of
∆a = 0.025 hsx
where:
hsx is the story height for level x. In this case, each level is 12 ft high; thus,
1. Determine the mass matrix [m] from the floor and roof seismic weight using details pre-
sented in Section 3.1.2 (ASCE-7, Section 12.7.2).
2. Determine the stiffness matrix [k] from the properties of the structural system using details
presented in Section 7.1.
3. With [m] and [k], solve the associated eigenvalue problem to determine modal properties
for each mode using details presented in Section 7.1.1:
a. Natural frequencies ωnj and periods Tnj using the eigenvalues ω 2j .
b. Mode shapes φij using normalized eigenvectors.
230 Introduction to Earthquake Engineering
c. Modal participation factors Γj (which are the contribution of each mode shape to the
total response) using Equation 7.50, repeated here for convenience,
∑
n
L miφij [Φ]T [ m]{1}
Γj = j = i=1
or {Γ} =
∑
n
Mj miφij2 [Φ]T [ m][Φ]
i=1
Where we have assumed that mii = mi since the mass matrix is diagonal.
d. Effective modal weight using Equation 7.66, repeated here for convenience,
2
∑
n
wiφij
∑
i=1 n
W je = or W je = Γ j wiφij or W je = Γ j L j g
∑
n
2 i=1
wφ
i ij
i=1
where:
wi = mi ⋅ g, which is the fraction of the building seismic weight at level i
n is the number of stories
4. Determine the number of modes to be used in the analysis (ASCE-7, Section 12.9.1).
Use a sufficient number of modes φij in each direction to account for at least 90% of the
actual weight W.
5. Develop the design response spectrum and determine the spectral accelerations for partici-
pating modes.
a. Develop an elastic response spectrum using details presented in Section 8.4.1, Figure
8.5 (ASCE-7, Section 11.4.5).
b. Determine spectral accelerations for the mth mode, Sam, using the elastic response
spectrum and the corresponding periods computed in step 3 for each contributing
mode found in step 4.
c. Divide Sam by the quantity R/Ie to obtain the modal seismic response coefficient,
Sam
Csm = (8.20)
R /I e
where:
R is the response modification factor in ASCE-7, Table 12.2-1
Ie is the importance factor in Table 8.2 (ASCE-7, Section 11.5.1)
where:
Vm is the seismic modal base shear computed in step 6a
Cvxm is the vertical distribution factor
wxφxm
Cvxm = (8.23)
∑
n
wiφim
i=1
where:
wx and wi are the weights at levels x and i, depicted in Figure 8.7
φim and φxm = ith and xth level displacements for the mth mode
We can then statistically combine these modal lateral forces using SRSS (Equation 7.68) or
CQC (Equations 7.69 and 7.70) to obtain the design lateral forces at each level Fx (ASCE-7,
Section 12.9.3).
8. Compute overturning moment. With the modal lateral forces obtained in the last step Fxm,
we can determine the modal overturning moments by applying static equilibrium,
M bm = ∑h F i im (8.24)
i=1
We can then statistically combine these modal overturning moments using SRSS (Equation
7.68) or CQC (Equations 7.69 and 7.70) to obtain the design overturning moment at the
base Mb (ASCE-7, Section 12.9.3).
9. Determine story drifts. Story drifts are the relative displacements of each story, computed
by taking the difference between displacements of two consecutive floors. Note that the
modal lateral loads used to calculate drifts are at the strength level (not service), as indicated
in ASCE-7, Figure 12.8-2.
a. The fraction of modal inelastic displacement contributed by the mth mode of vibration
at level x is
Cdδxem
δxm = (8.25)
Ie
232 Introduction to Earthquake Engineering
where:
Cd is the deflection amplification factor discussed in Section 8.3.2 and given in
ASCE-7, Table 12.2-1
Ie is the importance factor introduced in Section 8.4.2
δ xem is the deflection at level x in the mth mode, and can be determined using the
following relationship:
g T 2F
δxem = 2 m xm (8.26)
4π wx
where:
g is the acceleration due to gravity
Tm is the period of the mth mode obtained following step 3a
wx is the weight at level x
Fxm is the modal lateral forces determined in step 7
b. Statistically combine the modal inelastic displacements using SRSS (Equation 7.68) or
CQC (Equations 7.69 and 7.70) to obtain the design displacements.
c. The drift is then computed as
where:
δ xm is the deflection at the top of level x determined using Equation 8.25
δ xm−1 is the deflection at the bottom of level x determined using Equation 8.25
d. Statistically combine the modal drifts using SRSS (Equation 7.68) or CQC (Equations
7.69 and 7.70) to obtain the design drifts.
EXAMPLE 8.7
A four-story office building frame, with equal story heights of 12 ft, is located at a site with site soil
class D, SD1 = 0.75 and SDS = 1.33. The structural system is a special reinforced concrete moment
frame with the following stiffness matrix [k] and story weights of w1 = w2 = w3 = 789 kips and
w4 = 645 kips (Villaverde 2009). Use the MRS analysis to compute the design earthquake loads
acting on each floor, the story drifts, and the overturning moment:
Solution
1. Determine the mass matrix [m]. Use the floor and roof seismic weights given in the
problem statement (the matrix can be obtained using D’Alembert’s principle as shown
in Chapter 7, Example 8.4). The first floor is listed in the first row, while the roof is listed
along the last row (some books will reverse this order).
2.04 0 0 0
0 2.04 0 0
[m] = [w ]/g = kips ⋅ s2 /in
0 0 2.04 0
0 0 0 1.67
Seismic Code Provisions 233
where:
g = 386.4 in/s2
2. Determine the stiffness matrix [k]. Use the stiffness matrix given in the problem
statement (this matrix can be obtained using techniques from matrix structural analysis).
3. With [m] and [k], solve the associated eigenvalue problem to determine modal proper-
ties for each mode using MATLAB®:
a. Natural frequencies ωnj and periods Tnj using the eigenvalues ω 2j .
d. Calculate the effective modal weight using Equation 7.66, Wje = Γ j ∑ ni=1 w iφij .
∑
4
W1e = Γ1 w iφi1 = 1.3023(789(0.219 + 0.526 + 0.837) + 645(1)) = 2466 kips
i =1
Similarly,
W3e = 99 kips
W4e = 82 kips
4. Determine the number of modes to be used in the analysis. Use enough modes φij in
each direction to capture at least 90% of the actual weight W. The total weight of the
system is W = 4(789 kips) + 645 kips = 3012 kips. The percentages of the total weight
per mode are as follows:
Mode 1, W1e /W ×100 = 2466 kips/ 3012 kips = 81.8%
Mode 2, W2e /W ×100 = 365 kips/ 3012 kips = 12.1%
Mode 3, W3e /W ×100 = 99 kips/ 3012 kips = 3.3%
Mode 4, W4e /W ×100 = 82 kips/ 3012 kips = 2.7%
Therefore, we only need to include the first two modes, which capture 93.9% of W.
5. Develop the design response spectrum and determine spectral accelerations for partici-
pating modes.
a. Develop an elastic response spectrum using details presented in Section 8.4.1,
Figure 8.5.
Seismic Code Provisions 235
1.5
SDS = 1.33
1.25
T
Sa = SDS 0.4 + 0.6 = 0.53 = 7.25T
1 To
Sa (g)
SD1 0.75
0.5 Sa = =
T T
0.25
SD1 0.75
Ts = = = 0.56 s
SDS 1.33
0
0 0.5 1 1.5 2 2.5 3 3.5 4
To = 0.2Ts = 0.11 T (s)
b. Determine spectral accelerations using Figure E8.9 and the periods computed in
step 3 for each contributing mode found in step 4. Enter the graph with T1 and T2
(since only the first two modes are required) to acquire spectral accelerations for the
two participating modes:
T1 = 0.713 s > Ts = 0.56 s; therefore, Sa1 = SD1/T1 = 0.75/0.713 = 1.05
T2 = 0.241 s < Ts = 0.56 s; therefore, Sa2 = 1.33
c. Divide Sam by R/Ie to obtain the design modal seismic response coefficients, Csm,
Ie = 1.0 from Table 8.2, since an office building is classified with an Occupancy
Category II. R = 8, for a special moment reinforced concrete frame (case C5 in
ASCE-7, Table 12.2-1).
Thus,
Sa1 1.05
C s1 = = = 0.13
R /Ie 8 /1
Sa2 1.33
C s2 = = = 0.17
R /Ie 8 /1
6. Determine the system base shear, Vb.
a. Compute the seismic modal base shear for each of the participating modes with Csm
from step 5c and Wm from step 3d:
b. Combine V1 and V2 using SRSS (Equation 7.68) to obtain the base shear Vt = ΣVi 2 .
c. Verify that Vt is larger than a minimum base shear value, V, obtained with an ELFP
analysis described in Section 8.5. If Vt is less than 85% of V (with T = CuTa when
T > CuTa), then lateral force, story shears, and story moments determined using the
modal analysis should be multiplied by 0.85 V/Vt.
First, calculate the base shear V using the ELFP method, with T = CuTa:
Cu = 1.4 since SD1 > 0.4 (ASCE-7, Table 12.8-1)
Ta = Ct hnx as discussed in Section 8.5
where:
hn = 48 ft and is the building height
for a concrete moment-resisting frame in Table 3.2 (ASCE-7, Table 12.8-2)
Ct = 0.016
x = 0.9
So,
Ta = Ct hnx = 0.016(48 ft)0.9 = 0.53 s
Thus,
and
Furthermore,
V = 0.85(387 kips) = 329 kips > Vt = 326 kips
0.85
Vb = (Vt)(0.85 V/Vt) = 0.85 V = 329 kips
Fx1 = Cvx1V1
where:
V1 = 0.85 V/Vt (320 kips) = 1.01(320 kips) = 323 kips, computed in step 6a
w 4φ41 641kips(1)
F41 = (323 kips) = (323 kips) = 110 kips
1890 k ⋅ ft 1890 k ⋅ ft
Fx 2 = Cvx 2V2
where:
V2 = 0.85 V/Vt (62 kips) = 1.01(62 kips) = 62.6 kips, computed in step 6a
w 4φ42 641kips(1)
F42 = (62.6 kips) = (62.6 kips) = −48 kips
−826k ⋅ ft −826 k ⋅ ft
b. Combine lateral forces for mode 1, Fx1, and mode 2, Fx2, using SRSS (Equation 7.68)
to obtain the lateral forces, Fx = ΣFxi2 .
Mb 2 = ∑hF i i2 = 4112
( ) + 6124
( ) + 9.136
( ) − 48(48) = −20 k ⋅ ft
i =1
b. Combine Mb1 and Mb2 using SRSS (Equation 7.68) to obtain the system overturning
moment at the base Mb = ΣMbi2 .
Cd δxem
δxm =
Ie
where:
Cd = 5.5 for special moment reinforced concrete frames given in ASCE-7, Table
12.2-1
Ie = 1 for an office building
g T 2F
δxem = 2 m xm
4π w x
where:
Tm is the period of the mth mode obtained in step 3a
wx is the weight at level x from step 1
Fxm is the modal lateral forces obtained in step 7a without the factor 0.85(V/Vt)
b. Combine inelastic displacements using SRSS (Equation 7.68) to obtain total inelastic
displacements δx = Σδxi2 .
c. Determine drifts at each level using Equation 8.19. The story drifts associated with
mode 1 are determined below:
Similarly, the story drifts associated with mode 2 are determined as follows:
d. Combine drifts using SRSS (Equation 7.68) to obtain total drifts Δ x = ΣΔ2xi .
e. Check allowable story drifts per ASCE-7, Section 12.12.1 and Table 12.12-1. For
buildings in Occupancy Category II, this table gives an allowable drift (for four sto-
ries or less structures with interior walls, partitions, etc. designed to accommodate
story drifts) of
Δa = 0.025 hsx
where:
hsx is the story height for level x. In this case each level is 12 ft high; thus,
available. In order to include the effects of ductility and adjust the design response for various haz-
ard levels, the response parameters should be divided by R/Ie.
The mathematical model of the structure is analyzed by directly solving the coupled equations
of motion discussed in Chapter 7, Equation 7.45, but including damping,
PROBLEMS
8.1 A column of a steel-braced frame in a single-story building carries the following external
loading: dead load, PD = 35 kips, roof live load, PLr = 15 kips, and seismic load (from
horizontal effects), QE = ±15 kips. Also, SDS = 1.25, Ωo = 2.5, and ρ = 1.3. Determine the
maximum and minimum axial forces in the column using ASD load combinations. Repeat
the problem accounting for overstrength.
8.2 Given the following frame and loading (moment and axial load) shown, determine the
maximum and minimum design loads in column C5 using LRFD load combinations.
SDS = 0.65 and redundancy factor ρ = 1.0.
–20 k
F2
N3 N4
–50 k
F1
N2 N5 N6 D=0
L=0
QE = ±38 k-ft
C5
D = 18 kips
L=0
N7
QE = ±5 kips
N1 N8
242 Introduction to Earthquake Engineering
8.3 Given the following single-story office building frame and loading, determine the maxi-
mum and minimum design loads in the brace member A, and maximum design load in
column B using ASD load combinations. SDS = 0.65 and redundancy factor ρ = 1.3.
F
D=0
L=0
QE = ±33 kips
B
A
D = 50 k
L=0
QE = ±17 kips
8.4 Use ASCE-7, Table 12.2-1, to determine seismic force-resisting system parameters and
restrictions for a steel special concentrically braced frame (SSCBF). Also, list an advan-
tage and a disadvantage for SSCBF systems.
8.5 Given Ss = 1.82 and S1 = 0.68, and assuming a stiff soil profile, determine the site coef-
ficients Fa and Fv, the maximum spectral accelerations SMS and SM1, and the design spectral
accelerations SDS and SD1.
8.6 Determine the spectral response acceleration parameters and draw the response spec-
trum for the Golden 1 Center Arena site, located at 500 David J Stern Walk, Sacramento,
California. Assume that the site (soil) class is D.
8.7 Draw the design response spectrum for a site where there is very dense soil (Site Class C).
The geotechnical engineer has provided the mapped acceleration parameters (SS = 1.25
and S1 = 0.52) and the long-period TL = 16 s.
8.8 A seismic vulnerability assessment of a two-story ordinary building located in Stockton,
California, is being performed. If the building’s first story is 14 ft high and the second
story is 12 ft high, compute the approximate structural period Ta, the static design base
shear V, the lateral forces at each level Fx, and the overturning moment at the base. The
proposed structural system is a steel ordinary moment frame with an estimated weight of
50 kips on the first level and 20 kips on the roof. The geotechnical engineer has provided
the mapped design acceleration parameters (SDS = 0.672 and SD1 = 0.362).
8.9 Given a two-story, 25-foot average height, wood frame single-family residence located in
Stockton, California, compute the design base shear. The geotechnical engineer estimates
Seismic Code Provisions 243
that the house will be on Site Class D, with mapped acceleration parameters of SS = 1.75
and S1 = 0.90. The building in plan view is regular. The structural system being proposed
is a wood structural panel shearwall. The estimated weight on level 1 is 83 kips and level
2 is 52 kips.
8.10 Given a six-story office building frame with equal story heights of 12 ft located in
Brookside (Stockton, California), compute the design earthquake loads acting on each
floor. The building in plan view is square. The structural system being proposed is a steel
special concentrically braced frame with an estimated weight on each floor of 100 kips and
80 kips on the roof. The geotechnical engineering report has classified the site soil class
as D.
8.11 First, sketch the design response spectrum using SDS = 0.82 and SD1 = 0.47. Then, use
the equivalent lateral force procedure to compute the approximate structural period Ta,
the design base shear V, the lateral forces at each level Fx, and the overturning moment
at the base. The structural system for the office building is a steel ordinary moment frame.
w2 = 20 kips
12 ft
w1 = 50 kips
14 ft
C5
20 ft 20 ft
8.12 A four-story special moment-resisting steel frame with equal story heights of 15 ft has a
period of 0.7 s, occupancy category of III, and elastic displacement of 0.7 in. What is the
maximum inelastic displacement (δ x) for the first story?
8.13 A three-story hospital building with reinforced concrete special moment frames has floor
heights of h1 = 14 ft, h2 = 12 ft, and h3 = 10 ft. The elastic displacements at each story
are δ1e = 0.8 in, δ2e = 1.45 in, and δ3e = 2.15 in. SDS = 0.82 and SD1 = 0.47, SDC of D, and
period, T = 0.75 s. Find the design story drifts and check to ensure that they are within the
allowable drifts.
8.14 Given a four-story County Jail with 12-foot story heights and steel special concentrically
braced frames, determine design story drifts and verify that they are within allowable drift
limits. The equivalent lateral force procedure analysis results are shown below.
8.15 Use the equivalent lateral force procedure to compute the approximate structural period
Ta, the base shear V, the lateral forces at each level Fx, the overturning moment at the base,
and the inelastic story drifts given the elastic displacements shown. Also, determine if the
244 Introduction to Earthquake Engineering
drifts are within the allowable limits. The structural system is a special moment reinforced
concrete frame at a site with SDS = 0.82 and SD1 = 0.47.
400 kips
2.3 in
12 ft
500 kips
1.5 in
12 ft
500 kips
0.7 in
14 ft
8.16 A three-story office building with a special moment-resisting frame has floor heights
of h1 = 15 ft, h2 = 12 ft, and h3 = 12 ft. The building has an estimated weight on each
floor of 850 kips and 700 kips on the roof. SDS = 0.81, SD1 = 0.50 and seismic design cat-
egory D. A preliminary analysis yielded mode shapes and periods shown. First, use the
ELFP method to determine the seismic base shear, and then use the MRS analysis method
to determine effective weight for each mode, story forces for each mode, and the maximum
story shear forces by SRSS.
1.0 1.0
0.675 0.250
0.252 –0.345
T1 = 1.5 s T2 = 0.64 s
21 0 0
[λ] = 0 96.6 0 s−2
0 0 221.4
88.8
Assume {Sa} =
187.2 in/s
2
272.4
Seismic Code Provisions 245
m3 = 1 kip s2/in
k3 = 60 kips/in
m2 = 1.5 kip s2/in
k2 = 120 kips/in
m1 = 2 kip s2/in
k1 = 180 kips/in
8.18 Given the following eigensolution for a three-story building with floor weights of 1800 kips
for the first floor and roof, and 1200 kips for the second floor:
2.86 −0.657 0.387 15.1 377
[Φ] = 1.95 0.725 −1.61 and {ω} = 38.5 rad/s, which yield {Sa} = 346 in/s2
1.00
1.00 1.00 61.7 308
Determine the participation factors for each mode, effective weight for each mode, base
shears for each mode using the given design spectral accelerations, and the maximum base
shear using the SRSS method.
8.19 Given the following six-story office building frame, ground motion parameters, and mass
and stiffness matrices, use MATLAB to compute the displacements, design earthquake
loads acting at each floor, base shear, and overturning moment using the MRS analysis. Also,
compute the effective modal masses for the six modes and accumulated mass percentage.
Ground motion parameters:
Site Class = D
SDS = 0.323
SD1 = 0.186
Building frame parameters (steel intermediate moment frame):
Risk category = II (office building)
SDC = C
R = 4.5
Cd = 4
12 ft
12 ft
12 ft
12 ft
12 ft
12 ft
3.95 0 0 0 0 0
0 3.88 0 0 0 0
0 0 2.65 0 0 0
[ m] = kips ⋅ s2 /in
0 0 0 2.59 0 0
0 0 0 0 1.36 0
0 0 0 0 0 1.29
246 Introduction to Earthquake Engineering
REFERENCES
American Society of Civil Engineers, ASCE/SEI, Minimum Design Loads for Buildings and Other Structures,
Reston, VA, 2010.
Villaverde, R., Fundamental Concepts of Earthquake Engineering, CRC Press, Boca Raton, FL, 2009.
Index
A Critical damping coefficient, 68
Critically damped system, 67
Accelerations, 34, 37, 80 CSMIP, see California Strong Motion Instrumentation
Aftershocks, 20 Program
Allowable seismic force-resisting system, 206–207
Allowable stress design (ASD), 203–206
American Society of Civil Engineers (ASCE), 15 D
Ancash earthquake, 5 D’Alembert’s principle, 61, 65–66, 152, 171
ASCE-7, 42, 203, 215, 219 Damped single-degree-of-freedom system, 64; see also
earthquake loads based on ASCE-7 equivalent lateral Undamped single-degree-of-freedom system
force procedure, 220–229 free vibration response of damped systems, 65–71
ELFP analysis steps, 220–225 structural damping, 71–73
MRS analysis steps, 229 time-dependent forced damped vibration response,
ASCE, see American Society of Civil Engineers 73–78
ASD, see Allowable stress design time-dependent support accelerations, 78–82
Ashlar masonry, 15 Damped systems, free vibration response of, 65–71
Attenuation, 29 Damping, 32, 91
relations, 138 coefficient, 66
ratio, 71
B Dead loads (DL), 42, 204, 206, 221
Decoupling process, 176–177
Base shear Deflections, 157–159, 163–165, 182–187, 212
using participation factors and response spectra, Deformation response factor, 63, 74
157–159, 163–165, 182–187 Degree-of-freedom (DOF), 37–38, 171
by time-dependent forces and support excitations, Design respsponse spectrum, 213–218
83–85 Detailing, 211
Bending moment, 53–54 Deterministic approach, 34
Body waves, 21, 22 Diagonal matrix, 172
Braced frames, 51, 53–56 Dip-slip, 20
Bracketed duration, 30–31 Dip angle, 19
Direct integration methods, 114
C central difference method, 117–120
Newmark’s Beta method for linear systems, 120–121
California Strong Motion Instrumentation Program Nigam−Jennings algorithm, 114–117, 118
(CSMIP), 33 Discrete system, 152; see also Continuous system
Cantilever shear wall, 57 deflections, base shear, and moments, 157–159
Central difference method, 117–120 equation of motion, 152–156
Circular frequency, 42 example, 159–161
City of Los Angeles building code (1943), 16 generalized participation factor, frequency, and
Collapse earthquakes, 9 natural period, 157
comparison of building collapse and natural Displacements, 37
earthquakes, 10 DL, see Dead loads
Complementary solutions, 62–63, 74 DOF, see Degree-of-freedom
Complete quadratic combination methods Dry-stone walls, 15
(CQC methods), 169, 185 Ductility, 33, 71
Compliant systems, 33 Duhamel’s integral, 91, 95
Contemporary design spectral analysis, 138 Dynamic analysis, 37, 83
Continental lithosphere, 11 Dynamic magnification factor, see Deformation response
Continuous system, 165–167; see also Discrete system factor
deflections, base shear, and moments, 163–165
equation of motion, 162–163 E
generalized participation factor, frequency, and
natural period, 163 Earthquake, 1, 37
generalized SDOF, 162 accelerogram, 123
conv function, 102 Ancash earthquake, 5
Convolution integral, 91, 114 charactereristics, 29–31
CQC methods, see Complete quadratic combination development of mitigation strategies, 14–17
methods earthquake-resistant design, 14
247
248 Index
Great Peruvian earthquake, see Ancash earthquake script to solving eigenvalue, 189–190
Ground cracking, 3 Matrix methods, 169
Ground failures, 2–5 Maximum considered earthquake (MCE), 34, 213–214
Ground motion, response to, 111–114 Maximum design earthquake (MDE), 34, 215
Ground shaking, 7–8 Maximum probable earthquake (MPE), 35
MDOF system, see Multi-degree of freedom system
H Messina-Reggio earthquake, 15
Mexico City earthquake (1985), 31
Hanging wall, 20 Mitigation strategies development for seismic hazard
Harmonic excitation force, 61 effects, 14–17
Harmonic motion, 40 MM scale, see Modified Mercalli scale
Heat, 65 Modal analysis, 169
Horizontal equilibrium of FBD, 66 Modal mass matrices, 176
Hypocenter, 13, 19 Modal participation factor, 180
Modal response spectrum analysis, 229–240
I Modal stiffness matrices, 176
Mode shapes for MDOF system, 172, 177–179
IBC, see International Building Code decoupling process, 176–177
Implicit methods, 114 eigenvector, 174–175
Newmark’s Beta method for linear systems, 120–121 simple harmonic equation, 172–173
Impulse, 91 Modified Mercalli scale (MM scale), 24
impulsive force, 91 MM-I, 24
loading function, 91, 92 MM-XII, 24
response function, 94–95 Moment(s), 27, 28, 47, 57, 157–159, 163–165
response of SDOF system to, 91–93 using participation factors and response spectra, 182–187
Inelastic response spectrum, 140; see also Elastic MPE, see Maximum probable earthquake
response spectrum Multi-degree of freedom system (MDOF system), 1, 151,
design response spectrum, 222 169
elastoplastic response spectra for El Centro quake, 142 degrees of freedom, 171
example, 142–143 equations of motion, 171–187
inelastic design response spectrum, 143–146 response spectrum analysis method, 188–198
nonlinear equation of motion, 140–141 spectrum analysis, 2
Intensity of earthquake, 23–25 structural mode shapes, 170, 172–179
Internal shear force, 53–54 Multistory building systems, 152
International Building Code (IBC), 203, 207 Multistory shear buildings, 1
K N
Kanamori wave energy, 28 National Earthquake Hazards Reduction Program
(NEHRP), 32, 207
L Natural circular frequency, 39, 71
Natural cyclic frequency, 41
Lake Zone, 31 Natural frequency, see Natural circular frequency
Lateral force-resisting systems, 50 NEHRP, see National Earthquake Hazards Reduction
Left-lateral faults, 20 Program
Linear impulse and momentum, 92 Newmark’s Beta method for linear systems, 120–121
Liquefaction, 3 Newmark–Hall method, 1, 133, 134
Lithosphere, 11 Newton’s second law, 34, 91, 184
Live loads (L), 16, 204 Nigam−Jennings algorithm, 114–117, 118
Load and resistance factor design (LRFD), 203, 206 Nobi earthquake, 15
Local magnitude scale, see Richter scale Nonhomogeneous differential equations, 62, 80
Logarithmic decrement, 69 Nonlinear equation of motion, 140–141
Loma Prieta earthquake (1989), 3, 32 Nonperiodic forcing function, 91
Love waves, 22, 23 Normal faults, 20
LRFD, see Load and resistance factor design Normal modes, see eigenvector
M O
Magnitude of earthquake, see Earthquake—magnitude Oblique-slip, 20
Magnitude saturation, 27 Observation station, 19
Man-made earthquakes, 9–11 Occupancy categories, 219
Mass, 184 and importance factors, 220
MATrix LABoratory (MATLAB), 99–102, 171, 173 Occupancy Category II, 219
script to performing operation, 196–198 Oceanic lithosphere, 11
250 Index