You are on page 1of 52

DEVELOPMENT OF A JOINT SEISMIC HAZARD CURVE

FOR MULTIPLE SITE SEISMIC HAZARD


by
DARYN R HOBBS
B.S., University of Colorado Boulder, 2013
M.S., University of Colorado Boulder, 2013

A Masters Report submitted to the


Faculty of the Graduate School of the
University of Colorado in partial fulfillment
of the requirements for the degree of
Master of Science
Department of Civil, Environmental, and Architectural Engineering
2013

This Masters Report entitled:


Development of a Joint Seismic Hazard Curve for Multiple Site Seismic Hazard
written by Daryn Hobbs
has been approved for the
Department of Civil, Environmental, and Architectural Engineering

__________________________________
Professor Keith Porter, Ph.D., P.E.

__________________________________
Professor Ross Corotis, Ph.D., P.E.

__________________________________
Professor Abbie Liel, Ph.D., P.E.

Date______________

The final copy of this Masters Report has been examined by the signatories, and we
find that both the content and the form meet acceptable presentation standards
of scholarly work in the above mentioned discipline.

Hobbs, Daryn R (M.S., Civil Engineering)


Development of a Joint Seismic Hazard Curve for Multiple Site Seismic Hazard
Masters Report directed by Associate Research Professor Keith Porter
The state of the art in probabilistic seismic hazard analysis was examined
and these concepts, along with recent work in spatial ground motion correlation,
were extended to develop the notion of a joint seismic hazard curve: a 3-dimensional
surface that shows the probability of exceeding a ground motion intensity measure
at two sites within the same time period. Some elementary earthquake
characterization concepts were reviewed, i.e. local magnitude, moment magnitude,
etc., and different intensity measures were considered including macroseismic
intensity and instrumental intensity. Ultimately peak ground acceleration, a
measure of instrumental intensity, was chosen for use in the joint seismic hazard
curve although the same methodology could be used for spectral acceleration or
other measures of instrumental intensity. A seismic hazard curve was created using
data from the OpenSHA event set calculator, the theorem of total probability, and
the assumption of Poisson arrivals for earthquake occurrence. A methodology was
created for the creation of a joint seismic hazard curve for two sites with two
separate underlying assumptions: 1) Sites have conditionally independent ground
motions and 2) Sites have correlated ground motions. It was found that the
correlated ground motion assumption should not be used for sites farther than 10
km apart.

TABLE OF CONTENTS
1.

INTRODUCTION ......................................................................................................... 1
1.1. Problem Statement ................................................................................. 1
1.2. Objectives ................................................................................................ 2
1.3. Organization of Report ........................................................................... 2

2.

CHARACTERIZING EARTHQUAKES ............................................................................. 4


2.1. Magnitude ............................................................................................... 4
2.2. Macroseismic Intensity ........................................................................... 6
2.3. Instrumental Intensity ........................................................................... 7

3.

THE DEVELOPMENT OF SEISMIC HAZARD ............................................................... 10


3.1. Early Descriptions of Hazard ............................................................... 10
3.2. Evolution of PSHA ................................................................................ 12
3.2.1. Earthquake Rupture Forecasts ..................................................... 13
3.2.2. Ground Motion Prediction Equations............................................ 15
3.2.3. Probabilistic Calculation ................................................................ 20

4.

SPATIAL GROUND MOTION CORRELATION .............................................................. 23


4.1. Site Seismic Hazard and Separation Distance .................................... 24
4.2. Relationships for Correlation Coefficients ........................................... 27
4.3. Use of Semivariograms in Ground Motion Correlation ...................... 31

5.

JOINT SEISMIC HAZARD CURVES ............................................................................ 34


5.1. Conditionally Independent Joint Seismic Hazard Curve ................... 34
5.2. Correlated Joint Seismic Hazard Curve .............................................. 36

6.

CONCLUSIONS ........................................................................................................ 40

REFERENCES ................................................................................................................. 42

ii

LIST OF FIGURES
Figure 1 Seismic hazard curves..................................................................... 22
Figure 2 Test sites in San Francisco Bay Area, CA...................................... 24
Figure 3 Test sites in San Fernando Valley, CA .......................................... 25
Figure 4 Test sites in Memphis, TN .............................................................. 25
Figure 5 - Comparison of Correlation Coefficients .......................................... 30
Figure 6 Sample Semivariogram ................................................................... 32
Figure 7 Joint Seismic Hazard Curve, Same Site ........................................ 35
Figure 8 - Joint Seismic Hazard Curve, Separate Sites .................................. 36

iii

1. INTRODUCTION
1.1. PROBLEM STATEMENT
The concept of probabilistic seismic hazard dates back to the 1960s and has
since been continuously modified and refined (Field E. H., 2005). One of the
products of probabilistic seismic hazard analysis (PSHA) is a seismic hazard curve
that characterizes seismic hazard in terms of a ground motion intensity measure
(IM)Modified Mercalli Intensity (MMI), peak ground acceleration (PGA), damped
elastic spectral acceleration response (Sa) (henceforth all mention of Sa assumes 5%
damping), peak ground velocity (PGV), etc.for a single site in a finite period of
time, e.g. one year. The curve shows the probability of exceeding any probable value
of an IM in that time period. This curve is useful when planning for future seismic
risk and, in particular, it is used in the USGS National Seismic Hazard Mapping
Program. This information can then be transferred to engineers and policy makers
for use in the development of engineering standards and codes. To better
understand the seismic hazard at two or more sites the concept of a joint seismic
hazard curve is proposed. The joint curve considers the probability of exceeding an
IM at a site A and a site B at least once during the same earthquake in a specified
period of time.
A joint seismic hazard curve could be particularly useful in managing a group
of properties that vitally operate together to perform critical functions for an
organization or city. For example, consider two vehicular bridges that serve as the

only means of egress from an isolated community susceptible to earthquakes.


Knowing the probability that the ground motion will exceed a certain value at each
site at the same time would be valuable when designing or retrofitting the
structures, planning emergency evacuation routes, or for large earthquake
simulations such as ShakeOut. A joint seismic hazard curve would provide a
systems approach to understanding the seismic vulnerability of the communitys
major transportation infrastructure and could also be extended for use in
quantifying the risk of other natural hazards.

1.2. OBJECTIVES
This report first explores the framework necessary to create a seismic hazard
curve. It works within this framework to formulate the steps required to create a
joint seismic hazard curve for two sites under the assumption that their ground
motions are independent of each other. An attempt is then made to improve this
assumption by incorporating spatial ground motion correlation.

1.3. ORGANIZATION OF REPORT


This section introduces the problem and lays out the objectives of the report.
Section 2 provides a review of the methods used to characterize the size of an
earthquake and the resultant ground shaking. A detailed explanation of the PSHA
process is covered in Section 3, ending with a seismic hazard curve created by the
author. Section 4 covers recent work that has been completed on spatial ground
motion correlation and introduces empirical expressions for the correlation of the

ground motion at two sites as a function of distance. Section 5 explains the steps
taken to create a conditionally independent joint seismic hazard curve and a
correlated joint seismic hazard curve.

2. CHARACTERIZING EARTHQUAKES
2.1. MAGNITUDE
Several methods have been developed to quantify the amount of energy
released by an earthquake. Richters definition of magnitude is one of the most wellknown magnitude scales and was first developed in 1935. His scale uses data
describing

ground

motion

obtained

from

Wood-Anderson

seismograph;

specifically, the maximum amplitude from the horizontal ground motion


measurements taken 100 km away from the epicenter is used (Bullen & Bolt, 1985).
His equation is shown below:
(1)
Equation 1: Bullen & Bolt (1985)

where A (m) is the maximum amplitude traced on a seismogram located at a


distance

(km) from the epicenter and A0 is an empirical correction factor that

accounts for the distance the measurement was taken from the epicenter.
The magnitude found using this method is now referred to as the local
magnitude, ML. Instrumental limitations require this method to use data taken far
away from a very large earthquakes source and hence use data that are less
accurate in determining energy levels (Housner & Jennings, 1982). Further,
Richters scale is only appropriate for shallow earthquakes (Bullen & Bolt, 1985).
Refinement by Richter and Gutenberg resulted in a modified Richter scale.
Empirical tables were developed to calculate magnitudes for significantly deeper
earthquakes and two notions of magnitude were developed using measurements
4

from surface waves, Ms, and body waves, Mb. They determined that the two
measures were related and developed a relationship to calculate the energy
contained in these waves (Kanamori, 1978). This relationship is an approximation
as the scales only use surface and body waves with specific periods and do not
account for the entire spectrum of waves. Error exists when using this method to
estimate the energy of very large earthquakes that produce significant long-period
waves (Kanamori, 1978).
With advances in seismometer technology came advances in using longperiod wave data to describe the energy produced by very large earthquakes. The
notion of seismic moment, M0, was developed by Aki (1966) and can be calculated
with Equation 2:
(2)
Equation 2: Aki (1966)

where (dyn/cm2) is the rigidity of the material surrounding the fault, D (cm) is the
average displacement discontinuity along the rupture plane, and S (cm2) is the
surface area of the faults rupture. The seismic moment can then be used to
calculate the elastic strain energy released in an earthquake, shown in Equation 3:
(3)
Equation 3: Kanamori (1977)

where is the stress drop in the fault. Kanamori (1977) found that for large
earthquakes (Ms > 6.5) the stress drop is almost constant (see Equation 5).
Therefore by using the relationship for energy given in Equation 4, Equation 5 and

the energy-magnitude relationship developed by Gutenberg and Richter, shown in


Equation 6, the moment magnitude, Mw, can be calculated per Equation 7:

Equations 4: Kanamori (1977)

(4)
(5)
(6)
(7)

Equations 5-7: Kanamori (1978)

where M0 is in dyn-cm, H is the energy lost due to friction, and W is the wave
energy. Equation 5 neglects the heat loss during the fault rupture (Kanamori,
1977). Moment magnitude is the standard scale used today largely due to its lack of
saturation at large magnitudes.

2.2. MACROSEISMIC INTENSITY


An earthquakes effects on buildings, infrastructure, and people cannot
always be directly related to the amount of energy released, as described by its
magnitude. Intensity scales have been developed to better describe the relative
shaking intensity an earthquake produces and the subsequent damage. The
macroseismic parameters include those that can be described in laymans terms and
are a result of the general nature of the ground acceleration and the complex action
of the seismic waves. Since the above parameters are not easily quantified in a
holistic sense, a qualitative basis is used for intensity scales. These scales describe
macroseismic intensity and conglomerate the effects of the parameters mentioned
above to a simplified intensity level. The most prominent scale in North America is

the Modified Mercalli Intensity Scale (MMI). An earthquakes intensity is measured


on this scale by how damaging it was to different building types, how people reacted
to it, what the effects were on the geography of the surrounding area, and other
qualitative descriptions.
The MMI scale contains discrete levels ranging from I to XII. For example, at
level I no ground shaking is felt, at level VI the earthquake is felt by everyone but
there is no significant damage, at level IX heavy damage is caused with poorly
designed structures collapsing or experiencing severe damage, and at level XII
severe damage is incurred to all structures, surface waves can be seen on the
ground, and large ground masses are displaced or altered (Bullen & Bolt, 1985).
An intensity scale such as this can be used to map out regions that
experienced similar intensities which are separated by lines called isoseismal
curves (Bullen & Bolt, 1985). As waves travel farther from the source of the
earthquake, energy is dissipated and intensity decreases. Isoseismal lines look
similar to a topographical map with the largest intensity regions near the epicenter.

2.3. INSTRUMENTAL INTENSITY


Instrumental intensity measuresespecially PGA and Saare commonly
used to quantify ground shaking at specific locations during a particular size of
earthquake. In structural design, a nominal, maximum force is desired and this
must respect the dynamic nature of earthquake loads in combination with the
corresponding dynamic behavior of the structure.

The maximum ground

acceleration, and resultant structural response at a given site is of interest here,


7

and these are quantified by PGA and Sa. It is important to note that Housner and
Jennings (1982) suggest that using a single number can be problematic, It is
inherently impossible to describe a complex phenomenon by a single number, and a
great deal of information is inevitably lost when this is attempted. The nature of
the ground motion time history is also important; the effect of a given acceleration
can be different if it is short and abrupt than if it is long and powerful.
To calculate spectral acceleration, response spectra are created based on
historic earthquake data to account for the various damping characteristics and
modal frequencies that a structure may have under possible earthquake scenarios.
Response spectra are created using a time history of ground acceleration, velocity,
or displacement recorded during a single earthquake at a specific site. These data,
typically acceleration data, are then used as input for structural models that
calculate the dynamic response of a single degree-of-freedom elastic oscillator with a
certain natural period of vibration and damping ratio (most often 5%). The
calculations are repeated for many different natural periods and the maximum
relative displacement, relative velocity, and total acceleration is recorded for each
period. This is used to create a spectrum of the expected response to the specified
ground motion at a site. A response spectrum shows how an earthquake affects
structures

oscillating

at

different

frequencies

and

can

produce

spectral

accelerations, velocities, and displacements. This entire process can be repeated


using a different damping ratio each time, typically ranging from 0% to 10%
damping.

Response spectra are useful for retroactively determining how a structure


would respond to an earthquake at a given site, but a more involved approach must
be used for design purposes. Since it is nearly impossible to predict the precise
ground motion of an impending earthquake, even with historical data, design
spectra must consider the probability an earthquake will occur at all possible
locations that could affect the site, as well as the ground motion attenuation
between the earthquakes location and the site of interest. Further, to be inserted
into a design standard or code, the design spectra must be simplified and made
more conservative to be applicable on a regional basis. ASCE 7-10, for example,
offers a simplified design spectrum that has a constant design acceleration section
around Sa(0.2 sec, 5%) and a long period shape that represents the average of many
records. For these reasons, understanding site seismic hazard allows a more refined
analysis to be performed than what is required in a simplified building code.

3. THE DEVELOPMENT OF SEISMIC HAZARD


3.1. EARLY DESCRIPTIONS OF HAZARD
In the early years of modern seismology, late nineteenth and early twentieth
centuries, the concept of seismic hazard was ill-defined and poorly understood. It
was during this time that seismologists postulated that when the ground slips and
cracks along faults, earthquakes occur and not vice versa (Reitherman, 2012).
Discovering this relationship was an essential starting point in defining and
calculating seismic hazard. This crucial concept also provides a hint at the inherent
intricacies involved in determining hazard since fault sections and other potential
earthquake sources must be considered on a source-by-source basis.
Advancement can be seen in the post-earthquake publications of the 1906
San Francisco Earthquake and Fire. Serving as a major learning experience for
engineers and seismologists of the time, the analysis of the earthquakes destruction
resulted in many conclusions that would assist in the later developments of hazard
quantification. Some of these conclusions include the amplifying effect of soft and
alluvial soils, the notion of how complex ground motion attenuation can be, and the
ways in which a fault slip can propagate (Gilbert et al., 1907). Engineers realized
that these parameters were significant contributors to how damaging the San
Francisco Earthquake was to specific sites but were unable to develop distinct
relationships between them to be useful for assessing site seismic hazard in the
future. The conclusion that the damage at a site caused by an earthquake is a

10

function of the distance from the earthquake source and the type of ground the site
is on was vital in the development of seismic hazard.
One of the early methods for quantifying seismic hazard at a specific site was
developed by Cornell in 1968. He attempted to provide the method for
integrating the individual influences of potential earthquake sources, near and far,
more active or less, into the probability distribution of maximum annual intensity.
(Cornell, 1968). This is regarded as the first development of PSHA written in
English. One of the results of this approach is the number of expected earthquakes
in one year that will exceed a given level of MMI. More generally, this method can
produce the probability distribution of the greatest MMI that will occur at a site
throughout some interval of time (Cornell, 1968). Aside from MMI, Cornell
demonstrated that other variables of interest could be calculated if a relationship
existed between that variable and an earthquakes magnitude and focal distance.
Thus, in a similar fashion, calculations could be performed to create probability
distributions for instrumental intensity measures such as PGA, Sa, and PGV.
Cornells approach is an application of the theorem of total probability,
integrating over the occurrence frequency of nearby possible earthquake ruptures
and the resulting probabilistic ground motion at the site of interest, to arrive at an
exceedance rate (events per year, for example) of various levels of IM. If one
assumes that earthquakes can be treated as a Poisson process, one can then
estimate the probability that any specified value of IM will be exceeded during any
period of interest.

11

Cornells analytical approach relies heavily on its assumption that


earthquakes be modeled as Poisson arrivals. It assumes that earthquakes occur
independently of each other in time. As is pointed out in his paper, this assumption
does not consider aftershocks and is not consistent with elastic rebound theory. In
order for his method to be simple, it also considers that there is an equally likely
chance for an earthquake to occur everywhere along a fault. Cornells method
provided an analytical model of determining seismic hazard suitable for hand
calculations but a more detailed model including more refined empirical attenuation
relationships was necessary to increase its applicability.

3.2. EVOLUTION OF PSHA


Advances in PSHA have resulted in refined methods that address a breadth
of the variables that affect earthquake occurrence and subsequent ground motion.
Modern PSHA consists of three parts: one or more earthquake rupture forecasts
(each with an associated weight or Bayesian probability), one or more ground
motion prediction equations (GMPEs, again each with an associated weight or
Bayesian probability) and an application of the theorem of total probability similar
to that done by Cornell (1968) but with a logic tree to account for different
earthquake rupture forecasts and GMPEs (Field E. H., 2005). In the past three
decades there have been major advancements in earthquake rupture forecasts and
GMPEs which have provided a better grasp on seismic hazard, especially in the
Western United States.

12

3.2.1. EARTHQUAKE RUPTURE FORECASTS


As an attempt to address all factors contributing to earthquake occurrence,
some recent seismic earthquake rupture forecasts have become lengthy and
complex. These models use a brute-force method that exhaustively accounts for all
possible earthquake sources and all estimated sources of uncertainty. As time
passes with todays level of seismic recording technology, the amount of historical
data that researchers have available to them increases. These data serve as a
baseline for models to build off and a means to judge accuracy.
In recent years, major contributions to rupture forecasts have come from the
development of the Uniform California Earthquake Rupture Forecast (UCERF),
now in its second version developed in 2008 and referred to as UCERF 2. This
forecast addresses the first steps required in PSHA: to calculate a probability that a
given magnitude earthquake will occur in a region in a defined time span. UCERF 2
consists of four sets of models used in this order: fault models, deformation models,
earthquake rate models, and then probability models.
The fault models break all of Californias known faults into sections and
assign each section a set of parameters of interest such as geometry, average slip
rate, and aseismicity factor (Field et al., 2009). These sections are created and
assembled based on various theories and this results in two models: FM2.1 and
FM2.2. Next, deformation models use the fault models to calculate a slip rate for
each fault section. Six deformation models, DM2.1 through 2.3 for FM2.1 and
DM2.4 through 2.6 for FM2.2, were created to reduce epistemic uncertainty; the

13

main discrepancy being how to distribute the slip rates among neighboring faults.
Thirdly, Earthquake rate models take the above information and determine the rate
at which all possible damaging earthquakes could occur on each fault section. The
models created here are dependent on the type of source, type-A, -B, or -C, the
global slip rate between the North American and Pacific Plates, and the long-term
and short-term nature of how energy is released through earthquakes.
Type-A, -B, and -C sources are classified by the amount of information known
about the fault or region. Enough is known about type-A sources that permanent
rupture boundaries (segment endpoints) can be hypothesized and a stress-renewal
recurrence model can be applied (Field et al., 2009). Type-B sources include faults
that have slip-rate estimates but there are not enough historical data to create
stress-renewal probabilities. Type-C sources include area sources where there is not
enough information to assign slip to distinct faults.
One fault model reduces the rate at which earthquakes occur in order to
produce larger events while another does the opposite. A segmented model only
allows earthquakes to occur along predefined segments of fault sections, disallowing
fault-to-fault ruptures and an unsegmented model allows for fault ruptures to jump
between segments. Finally, probability models describe how the earthquakes could
occur throughout a given period of time. The probability models can be grouped into
two categories: time independent and time dependent (Field et al., 2009).
One of the defining characteristics of UCERF 2 is its means of managing
epistemic uncertainties by way of a logic tree. Each of the previously mentioned sets

14

of models represents the seismological communitys best guess as to how the


phenomenon should be modeled and with which variables. At this time, it is not
known which model perfectly predicts the phenomenon of interest. To manage these
uncertainties the logic tree considers the leading models and weights them
according to the UCERF 2 authors degree of belief in each model. However, this
tree has 480 branches that have varying results and for this reason efforts have
been made to trim the tree and use only the uncertainties that contribute the most
(Porter et al., 2012). The full logic tree is somewhat computationally intensive and
in the end, an average value is used for the final probability of earthquake
occurrence (Field et al., 2009).
3.2.2. GROUND MOTION P REDICTION EQUATIONS
Earthquake rupture forecasts contain the information needed to quantify the
seismicity of a region and the next step in PSHA is to use GMPEs to determine how
strongly an earthquake will shake a particular site within that region. As used
previously by Cornell, the two main input variables among the early GMPEs are
earthquake magnitude and source-to-site distance.
Early attenuation calculations were theoretically based and followed physicsbased attenuation law. They used the exponential decay function with the following
independent variables: wave frequency, travel path, velocity, and an attenuation
constant which was dependent on geologic conditions (Trifunac & Brady, 1975).
Trifunac and Brady also mention the use of finite element analysis but conclude
that this method is not ideal because it is too computationally expensive and there

15

are a lack of available data required to create a precise geophysical ground model.
There was a trend in the 1980s and 1990s of researchers not only working on
mathematically based analytical models, but also producing empirical ones.
Different approaches were taken in curve-fitting and multiple equations have been
developed with this methodology.
Similar to the physics-based attenuation law, the functional form of most
regression-based equations is exponential; one reason being that the definition of
moment magnitude is exponential (Boore & Joyner, 1982). These equations started
out with separate terms for magnitude, distance, and site characteristics.
Regression coefficients are on most or all terms. Equations 8-10 are examples of
early GMPEs with variable names modified for ease of comparison:

Equation 8: Joyner & Boore (1982)


Equation 9: Campbell (1988)

(8)
(9)
(10)

Equation 10: Boore Joyner, & Fumal (1993)

where in all equations Y is the IM of interest: peak ground acceleration, velocity,


displacement, or spectral acceleration, M is moment magnitude, and r is some
measure of source-to-site distance. In Equation 8, S accounts for site soil conditions
and is binary: 0 for rock and 1 for soil, and P represents the uncertainty in the
prediction. In Equation 9, s is a function of fault type, directivity, soil type, building
size, and building embedment. In Equation 10, GB and GC represent site
classification and are binary: GB=1 for site class B and GC=1 for site class C and
both are zero otherwise, r and e are variables that account for the variability
16

within each earthquake record and between earthquakes, respectively. All other
variables are regression coefficients and can be found in tables provided by the
appropriate authors. Site class is defined by VS30 where for site class A VS30 is
greater than 750 m/s, for site class B VS30 is less than 750 m/s and greater than 360
m/s, for site class C VS30 is less than 360 m/s and greater than 180 m/s, and for site
class D VS30 is less than 180 m/s. Equation 10 is not applicable for site class D.
The above equations use different measures of source-to-site distance. This
includes hypocentral distance, epicentral distance, and various distances to the
rupture surface. Some variables, including distance, are magnitude dependent.
Equation 9 uses either surface wave magnitude or local magnitude but Equations 8
and 10 use moment magnitude. In later research, moment magnitude prevails and
is prominent throughout all modern equations. Also seen in Equations 8-10 are
differing methods of site classification. Equation 9 uses a function, s, to account for
site classification, faulting parameters and even building attributes while
Equations 8 and 10 use simple binary switches that turn regression coefficients on
or off.
In 1997, Boore, Joyner and Fumal as well as Abrahamson and Silva refined
previous regression models and used more sophisticated site classification variables.
Their models are shown in brief in Equations 11 and 12:
(11)
(

(12)

Equation 11: Boore Joyner, & Fumal (1997), Equation 12: Abrahamson & Silva (1997)

17

where in both equations M is moment magnitude. In Equation 11 Vs is the average


shear-wave velocity in the top 30 m of earth, r is a function of epicentral distance
and a fictitious depth, and the bis and VA are determined via regression. In
Equation 12 the fis represent functions of the enclosed variables and each function
contains its own regression coefficients, rrup is the closest distance to the rupture
plane, F accounts for fault type, HW accounts for the hanging wall effect, S accounts
for site class, and pgarock is the expected peak ground acceleration on rock, and is
used to estimate non-linear soil response.
Over time, as GMPEs used terms for more source, site, and path variables
and the data used in the regression became more numerous and included larger
magnitude earthquakes, the applicability of the equations increased. This is true for
both magnitude and distance. For example, the most recent equations initiated by
the Next Generation Attenuation (NGA) Program can be applied to earthquakes
with magnitudes between 4 and 8.5 and at distances from 0-200 km whereas an
early GMPE, i.e. Boore, Joyner, and Fumal (1993), could only be applied to
magnitudes between 5 and 7.7 and at distances below 100 km. Two of the NGA
equations are shown below:
(

(13)

Equation 13: Boore & Atkinson (2008)

(
(

)
)

(14)

Equation 14: Abrahamson & Silva (2008)

where in Equation 13 FM, FD, and FS are magnitude, distance, and site
amplification functions, respectively, Rjb is the Joyner-Boore distance which is the

18

closest horizontal distance to the surface projection of the fault, is the fractional
number of standard deviations of a single predicted value of ln(Y) away from the
mean of ln(Y), and represents the inter- and intra-event aleatory uncertainty
(Boore & Atkinson, 2008). In Equation 14 the fis are functions of the enclosed
variables, FRV, FNM, FAS, and FHW, are flags for reverse faulting, normal faulting,
aftershocks, and hanging wall effects, respectively, PGA1100 is the median peak
acceleration for VS30 = 1100 m/s, Rx is the horizontal distance from the top edge of
rupture measured perpendicular to the fault strike, is the fault dip angle, ZTOR is
the depth to top of fault rupture, and Z1.0 is the depth to VS=1.0 km/s (Abrahamson
& Silva, 2008).
The NGA equations have an increased range of applicability over the
previous GMPEs but not without added complexity. Site classification was
quantified using a continuous variable, Vs30, and the type of faulting was described
in more detail. One may have to go through as many as six equations to calculate
the soil depth term, f10. However, the NGA equation authors suggest an increase in
applicability over previous GMPEs and even extrapolated beyond earthquake
magnitudes used in the regression analysis. The availability of strong-motion data
obtained close to the source (R < 20 km) of large earthquakes (M > ~7.0) has limited
GMPEs (Joyner & Boore, 1988) but as more data become available over time, it is
likely that new equations have potential for higher accuracy and wider
applicability.

19

3.2.3. PROBABILISTIC CALCULATION


The last step in PSHA is to combine earthquake rupture forecasts and
GMPEs mentioned above, calculating the level of hazard on a site-specific level. The
earthquake rupture forecasts provide the rate at which earthquakes of various,
discrete magnitudes occur throughout time. A GMPE takes certain characteristics
of each of these earthquakesmagnitude, distance from site, fault type, etc.along
with the site conditions to calculate the probability distribution for the desired IM:
PGA, Sa, PGV, etc. For engineering design purposes, it is the probability that an IM
will be exceeded that is of greatest interest. The rate

that a ground motion will be

exceeded in time span T is calculated by multiplying the rate, , at which any


earthquake i will occur by the probability that an IM will exceed a certain value
given that earthquake occurs, then summing this product over all possible
earthquakes that can cause substantial ground shaking at the site, as shown in
Equation 15:

where Ei represents any one of N earthquakes that occurs with rate

(15)
, IMA is the

intensity measure at site A, and IM1 is some arbitrary intensity measure. It is


important to note that one should avoid including earthquakes with a source-to-site
distance greater than the applicable distance of the GMPE being used and with a
magnitude outside of the GMPEs applicable range.

20

To arrive at a probability of exceedance, it is convenient to assume that


earthquakes occur as Poisson events. Equation 15 is used in Equation 16, to
calculate the probability of at least one occurrence of IM1 being exceeded, resulting
in Equation 17.
(16)
(

(17)

This process can be repeated to create a seismic hazard curve for any site in a
region sufficiently described by an earthquake rupture forecast and with applicable
GMPEs.
An example of a seismic hazard curve was created using the steps described
above and can be seen in Figure 1. The Event Set Calculator from OpenSHA was
used to collect the required data and perform the calculations needed to find the
effective earthquake sources and their respective rates as well as the probability
distributions of the ground shaking at the site; both terms are shown in Equation
15. UCERF 2 was used to identify earthquake sources and their respective rates of
occurrence at varying magnitudes and the Boore & Atkinson (2008) GMPE was
used to calculate mean PGA at sites in the San Francsico Bay Area. Then Equation
10 was used to calculate the probabilities of exceedance for various PGA levels.

21

Figure 1 Seismic hazard curves shown for two sites 7.2 km apart. Site B is base of the new San
Francisco-Oakland Bay Bridge tower (374854N, 1222132W) and Site A is 7.2 km northeast of Site B
(375030N, 122173W).

22

4. SPATIAL GROUND MOTION CORRELATION


There are situations where one must quantify the ground shaking at
multiple, closely spaced sites instead of studying ground shaking on an individual
site basis. In doing so it is important to consider spatial ground motion correlation,
or the tendency of two sites separated in space to have similar ground shaking
under the same earthquake. For example, in a series of earthquakes two sites may
have PGAs that both tend to be above or below the mean PGA, calculated with a
GMPE, in each earthquake. Physically, this correlation can be attributed to the
similar travel paths of seismic waves as well as both sites being a similar distance
from discontinuities along the fault if the fault rupture is very long compared to the
source-to-site distance (Park et al., 2007).
Including ground motion correlation at multiple sites is of considerable
importance in the risk assessment of a portfolio of properties in close proximity
(Park et al., 2007; Jayaram & Baker, 2009). Park et al. (2007) note that by
accounting for the correlation of ground motion between multiple sites, portfolio loss
estimates become more accurate; if spatial ground motion correlation is not
accounted for then the occurrence frequency of very strong ground motion at both
sites tends to be underestimated and the occurrence frequency of weaker ground
motion at both sites tends to be overestimated. Further, utilizing correlation can be
of particular interest to increase the accuracy of loss estimates and ground shaking
maps that are produced immediately after an earthquake occursless than 20
minutes after an event (David Wald, personal communication, March 15, 2013)

23

such as the USGS PAGER and ShakeMaps programs (Park et al., 2007). These
prompt post-earthquake reports are especially helpful in an emergency response
context where response crews need to be dispatched to the most severely affected
areas first.

4.1. SITE SEISMIC HAZARD AND SEPARATION DISTANCE


The author performed an investigation studying how the seismic hazard
between two sites varies with distance for three regions: San Francisco Bay Area,
CA; San Fernando Valley, CA; and Memphis, TN. The purpose of this investigation
was to gain a sense of the distance scale for which seismic hazard differs
significantly between two sites as a function of separation distance. Might sites
have similar seismic hazard if they are 100 km apart? 10 km? 1 km? A main site of
interest was chosen and several sites surrounding the main site located along two
azimuths and at various distances. The azimuths were chosen to be parallel and
perpendicular to faults immediately surrounding the site. See Figures 2-4.

Figure 2 Test sites in San Francisco Bay Area, CA. Faults shown in red.

24

Figure 3 Test sites in San Fernando Valley, CA. Faults shown in red.

Figure 4 Test sites in Memphis, TN.

Then a seismic hazard curve was created for each site using the USGS Java
Ground Motion Parameter Calculator. This tool uses USGS data from 2002 to
calculate seismic hazard assuming site conditions are on the NEHRP Site Class B-C
boundary where Vs30 = 760 m/s. Specific probabilities of exceedance (PE) were
examined for comparison including 50% in 10 years, 10% in 50 years, 5% in 50
years, and 2% in 50 years.

25

The results, shown in Table 1, show that for sites chosen perpendicular to the
surrounding faults, as distance from the main site increases, similarity to the PGA
of the main site decreases at a faster rate than for sites chosen parallel to the
surrounding faults. For example when the new Bay Bridge tower in the San
Francisco Bay Area was the main site, a second site along a parallel azimuth to
nearby faults could be as far as 40 km away from the main site before PGA (2% PE
in 50 years) differed by more than 10% from the main site while a second site only 4
km away along a perpendicular azimuth differed by more than 10%. It was
concluded that on a particular site basis there are factors other than separation
distance that affect the similarity of one sites seismic hazard to anothers. These
include directivity in relation to faults and soil conditions. Although they were not
considered in this particular investigation, it can be expected that site soil
conditions can significantly increase the difference in PGA between two sites
considering all modern GMPEs account for soil conditions.
Table 1 - Results for 10% PE in 50 Years

Separation Difference in PGA


Distance (km) from Main Site

Location

Directionality

San Francisco
Bay Area

Nearby Faults

7.2

1%

Nearby Faults

7.2

39%

San Fernando
Valley

Nearby Faults

11.7

17%

Nearby Faults

12.9

36%

26

4.2. RELATIONSHIPS FOR CORRELATION COEFFICIENTS


Formal studies have been completed on spatial correlation of ground motions
within a single earthquake in the past two decades. These start with the general
formulation of a GMPE shown in Equation 18:

(18)

where is the median ground motion in an earthquake i at site j,


the inter- and intra-event residuals respectively, and

and

and

are

are the standard

deviations for the respective residuals. This general form is used by many recent
authors who separate inter- and intra-event uncertainty.
Next, a spatial correlation coefficient is introduced to quantify the correlation
between sites and thus modify the

term (Park et al., 2007). In essence this

process adjusts the predicted ground motion parameter away from the mean to
account for the correlation that exists between the site of interest and a nearby site.
The correlation coefficient, , is defined in Equation 19:
(19)

where

is the variance of the differences of the natural logarithm of the motion

at two sites or the unexplained variance, and

is the variance of the natural

logarithm of the motion at a single site or the total variance.


Multiple equations have been created to calculate using different data sets.
These equations must satisfy the condition that at inter-site distances of 0 and

the correlation coefficient must be 1 and 0, respectively. Thus, most equations follow
27

the form of exponential decay (Abrahamson & Sykora, 1993; Boore et al., 2003; Park
et al., 2007; Goda & Hong, 2008; and Goda & Atkinson, 2009) and this general form
for PGA and SA(TN) is shown in Equation 20. Equations 21-22 include independent
variables that are frequency dependent indicating that correlation tends to increase
for longer periods and decrease for shorter periods (Abrahamson & Sykora, 1993):
(

(20)

Equation 20: Goda & Hong (2008)

Equation 21: Goda & Atkinson (2009)

Equation 22: Abrahamson & Sykora (1993)

(21)
(22)

from Boore & Atkinson (2008)

(23)

where , , c1, and c2 are constants, and is the separation distance between two
sites in km. Equation 21 has been modified from its original form to extract a
comparable correlation coefficient by assuming the value of

to be the intra-

event standard deviation calculated with the NGA GMPE from Boore & Atkinson
(2008).
Equations 24-28 are examples of correlation coefficient relationships that are
not frequency dependent and Park et al. (2007) examine the effects of simplifying
Equation 20 to Equations 27 and 28:
(

(24)

)
(

(25)

28

))

(26)

Equations 24-26: Boore et al. (2003)

Equations 27 & 28 : Park et al. 2007

(27)
(28)

Importantly restated, the equations above were established using different


data setse.g. Goda & Atkinson (2009) use data from the K-NET and KiK-net
networks in Japan while Boore et al. (2003) use data only from the 1994 Northridge
Earthquake; both studies examine PGA and Saand their differences imply that
the decay of spatial correlation can depend on geographic region and type of
earthquake (Goda & Atkinson, 2009). It was not pertinent for this report to perform
a detailed analysis comparing the equations shown above but a graphical
comparison is provided in Figure 5 to show their variability. This report is most
concerned with the existence of these equations and how they can be incorporated
into joint seismic hazard.
Although the above equations were developed for PGA and Sa, correlation
coefficients also exist in terms of PGV. Wang & Takada (2005) perform a study on
PGV correlation and adopt the same exponential decay expression shown in
Equation 20. Their results show a separation distance range of 20-40 km before
ground motions are completely uncorrelated.

29

Abrahamson & Sykora (1993)


Boore et al. (2003)
Park et al. (2007) (Eq. 26)
Park et al. (2007) (Eq. 27)
Goda & Hong (2008)
Goda & Atkinson (2009)

Correlation Coefficient,

0.8

0.6

0.4

0.2

0
0

10

Separation Distance, (km)


Figure 5 - Comparison of Equations 10-18 showing their variability
and how correlation diminishes quickly with separation distance.

The correlation coefficients drop off very quickly after a few kilometers of
separation distance between sites, notably in Equations 20 and 28. Equations 20,
21, 27, and 28 show a general lack of correlation around = 10 km in Figure 5. In
these studies the correlation of ground motions in a single earthquake were of
interest while in the authors investigation in Section 4.1, the similarity in overall
seismic hazard as a function of separation distance was of interest. Although this
difference exists, preventing the studies from being directly comparable to each
other, it is important to note that the lack of correlation around = 10 km
corresponds well with the results of the investigation in Section 4.1.

30

4.3. USE OF SEMIVARIOGRAMS IN GROUND MOTION CORRELATION


Another way spatial correlation has been described in recent work is by use
of semivariograms. This technique is typically used in geostatistics and provides a
measure of how dissimilar two random variables can be in relation to distance; put
simply, it quantifies the assumption that the closer two sites are, the more alike
their ground motions will be in the same earthquake. The functional form is shown
in Equation 29:
{

(29)

Equation 29: Jayaram & Baker (2009)

where
and

is the semivariogram function,

is the spectral acceleration at site u,

is the spectral acceleration at a site a distance

from site u. Jayaram and

Baker (2009) show that the semivariograms created using PGA and Sa data from
the 1994 Northridge and 1999 Chi-Chi Earthquakes are isotropic and hence, need
not incorporate the directionality between sites.
Two important parameters produced by a semivariogram are the range and
sill. The range is the distance at which the ground motion is no longer correlated.
The peak y-value of the semivariogram, realized at the range, is called the sill. If a
sill exists then this implies there is a distance where ground motion is no longer
correlated (Chiles & Delfiner, 2012). Jayaram & Baker (2009) show the relationship
between a semivariogram and the correlation coefficient in Equation 30:

(30)

Equation 30: Jayaram & Baker (2009)

31

where

is the experimental stationary semivariogram which is estimated from a

sample of data and

is an estimate of the ground motion correlation coefficient,

. Equation 30 uses normalized intra-event residuals so that

has a sill of 1

implying that there exists a separation distance where ground motions are
uncorrelated. For more on the development of Equation 30 see Jayaram & Baker
(2009). A sample semivariogram for visualization purposes is shown in Figure 6
using Equation 31 from Jayaram & Baker (2009).

(31)

Equation 31: Jayaram & Baker (2009)

0.8

( )

0.6

0.4

0.2

0
0

10

20

30

40

50

Separation Distance, (km)

Figure 6 Sample Semivariogram from Equation 31

Jayaram & Baker (2009) confirm the results of the frequency dependent
correlation coefficients described in Section 4.2: that correlation tends to increase

32

with period. This is shown as an increasing range in the semivariograms calculated


for the Northridge and Chi-Chi earthquakes when range was plotted against period.

33

5. JOINT SEISMIC HAZARD CURVES


Let a joint seismic hazard curve refer to the combining of hazard curves for
two sites into one 3-dimensional surface where the IM at Sites A and B are along
the horizontal x- and y-axes and the probability of exceeding both of those IMs on
the vertical z-axis. Rather than just understanding probabilistic seismic hazard at a
single site, a joint seismic hazard curve provides a better understanding of hazard
at two different sites that vitally act together.

5.1. CONDITIONALLY INDEPENDENT JOINT SEISMIC HAZARD CURVE


It is proposed here that the form of a conditionally independent joint seismic
hazard curve takes on a 3-D surface with IM at sites A and B along the x- and yaxes and PE on the vertical z-axis. The x-z plane where y=0 and y-z plane where
x=0 show the independent hazard curves for each site separately. The calculation of
the Probabilities of exceedance for the rest of the x-y plane, away from the axes,
starts with the equation for the joint probability of two independent random
variables shown in Equation 32:
|

(32)

where IMA and IMB are intensity measures at sites A and B, respectively and IM1
and IM2 are some arbitrary values to be exceeded. This assumes that IMA and IMB
are independent random variables. Next Equation 32 is inserted in Equation 15
and followed through to Equation 17 to arrive at Equation 33:

34

(33)

To develop the conditionally independent joint seismic hazard curve the same
procedure described in Section 3.2.3 was used except with Equation 33 in lieu of
Equation 17. Two curves were created: Figure 7 shows perfect correlation, i.e. Site A
and Site B are the same site, and Figure 8 is for two sites that are 7.2 km apart.

Figure 7 Joint Seismic Hazard Curve showing PGA with 50-year probability of exceedance
for the base of the new San Francisco-Oakland Bay Bridge tower (374854N, 1222132W).
Sites A and B are perfectly correlated, i.e. Site A=Site B.

35

Figure 8 - Joint Seismic Hazard Curve showing PGA with 50-year probability of exceedance.
Site B is the base of the new San Francisco-Oakland Bay Bridge tower (374854N, 1222132W)
and Site A is 7.2 km northeast of Site B (375030N, 122173W).

As stated previously, the curves, or surfaces rather, show the 2-dimensional


hazard curves for sites A and B on the x-z plane where IMB=0 and y-z plane where
IMA=0, respectively. Figure 7 shows that for a given IM at Site A, as the IM at Site
B increases, the PE stays constant until IMA = IMB, and vice versa. Figure 8 does
not show this same effect and the probability of exceeding the same IM at both sites
tends to be less than in Figure 7. This can be attributed to the calculation of the
joint probability using Equation 32.

5.2. CORRELATED JOINT SEISMIC HAZARD CURVE


If ground motion correlation is accounted for, then the covariance of IMA and
IMB conditioned on event Ei is no longer zero and they cannot be treated as
conditionally independent events. In this case, Equation 34 must be used:
|

(34)

36

where the exceedance probability of an IM at one site is influenced by the IM at


another site. This conditional probability allows for treating correlation between the
two sites in the following manner. Let

and

represent the expected value of the

intensity measure at sites A and B, respectively. The exceedance probability of IMA


(denoted by the second term on the right-hand side of Equation 34) is written out in
Equation 35 using the cumulative distribution function.
|

(35)

Following the procedure used by Boore et al. (2003), the natural logarithm of
IM at site A is assigned as the expected value of the natural logarithm at site B,
shown in Equation 36:
(36)
Next the first term on the right-hand side of Equation 34 is defined in
Equation 37. The numerator represents the conditional exceedance probability of
IMA and IMB. In the procedure used by Boore et al. (2003), the exceedance
probability of IMB depends on each distinct value of IMA and thus the numerator is
necessarily an integral, rather than a discrete summation. Note that the second
term in the integral uses the probability density function, not to be confused with
the cumulative distribution function used in the first term. Since only the
conditional exceedance probability of IMB given IMA is of interest, the exceedance
probability of IMA must be divided out and exist in the denominator.
|

|
|
37

))

(37)

(
where

and

))

are the median and logarithmic standard deviation of

IMB conditioned on IMA = x, and thus accounting for correlation. Now, taking
advantage of how Boore et al. (2003) take
Equation 24 to substitute

for

(shown in Equation 36), using

and canceling

in the

numerator and denominator, Equation 38 is derived:


|
(
where

))

(38)

. The author made no attempt to find a closed-form

solution to Equation 38. If a closed-form solution does not exist, it should be solved
numerically. Note that the range of applicability of Equation 38 should be no
greater than 10 km as Equation 24 is best applied for separation distances less than
10 km (Boore et al., 2003).
The same steps used in formulating the probability of exceedance considering
all earthquakes is used for the correlated joint seismic hazard curve and is shown in
Equation 39.

(39)

38

To develop a correlated joint seismic hazard curve, the same procedure


described in Section 3.2.3 should be used with Equation 39. Creating a numerical
solution to Equation 39 was beyond the scope of this report and thus a correlated
joint seismic hazard curve was not created.
A recommended procedure has been developed for the creation of a joint
seismic hazard curve. First the curves should be created for sites within regions no
larger than 10 km. A relationship for the correlation coefficient should be
determined based on earthquake records from this region by fitting the data to a
semivariogram, described in Section 4.3. The semivariogram should be examined
closely at the separation distances between the sites of interest. Then the
correlation coefficient should be used to relate the ground motion at one site to the
ground motion at the other (Equation 38) and arrive at a new rate of exceedance to
be used in the traditional Poisson distribution equation (Equation 39).

39

6. CONCLUSIONS
The state of the art in PSHA was explored so that the same principles could
be applied in creating a joint seismic hazard curve. An investigation into spatial
ground motion correlation was completed and compared with recent work to
conclude that beyond a separation distance of 10 km, incorporating correlation is
not useful. OpenSHA tools were used to identify earthquake sources and assign
rates of occurrence from UCERF 2 and subsequently calculate median PGA at Sites
A and B in the San Francisco Bay Area for each earthquake source. Earthquakes
were modeled as Poisson events and a probability of exceedance was calculated for
many PGA values considering all of the earthquake sources. This was carried out
using two assumptions: 1) the ground motion at Site A is independent of the ground
motion at Site B and 2) the ground motion at Site B was dependent on the ground
motion at Site A, incorporating spatial ground motion correlation. If spatial ground
motion correlation is not accounted for, the occurrence frequency of strong ground
motion at both sites tends to be underestimated and the occurrence frequency of
weaker ground motion at both sites tends to be overestimated. The range of
applicability of a joint seismic hazard curve is limited by the separation distance
between two sites.
Future research should be completed on the regional variability of ground
motion correlationmost of the work discussed in this report was completed in
California or Japanas well as the effect of other variables aside from separation
distance including soil conditions and directivity. Also, further work should be

40

completed to develop a closed form expression for Equation 39 and to solve it


numerically so that a correlated joint seismic hazard curve can be created. Lastly
the concepts covered in this report regarding the theorem of total probability,
empirical prediction equations, and spatial correlation could be applied to the risk
assessment of other natural hazards aside from earthquakes.

41

REFERENCES
Abrahamson, N., & Silva, W. (1997). Empirical response spectral attenuation
relations for shallow crustal earthquakes. Seismological Research Letters,
68(1), 94-127.
Abrahamson, N., & Silva, W. (2008). Summary of the Abrahamson & Silva NGA
Ground-Motion Relations. Earthquake Spectra, 24(1), 67-97.
Abrahamson, N., & Sykora, D. (1993). Variation of ground motions across individual
sites. Fourth DOE Natural Phenomena Hazards Mitigation Conference , 192198.
Aki, K. (1966). Generation and propagation of G waves from the Niigata
Earthquake of June 16, 1964. Bulletin of the Earthquake Research Institute,
44, 73-88.
Ang, A., & Tang, W. (2007). Probability concepts in engineering. Hoboken, NJ: John
Wiley & Sons, Inc.
Archuleta, R. J., Joyner, W. B., & Boore, D. M. (1979). A methodology for predicting
ground motion at specific sites. (E. Brabb, Ed.) Progress on Seismic Zonation
in the San Francisco Bay Region, 26-36.
Atkinson, G. M., & Silva, W. (2000). Stochastic modeling of california ground
motions. Bulletin of the Seismological Society of America, 255-274.
Boore, D. M., & Atkinson, G. M. (2008). Ground-motion prediction equations for the
average horizontal component of PGA, PGV, and 5%-damped PSA at spectral
periods between 0.01 s and 10.0 s. Earthquake Spectra, 99-138.
Boore, D. M., & Joyner, W. B. (1982). The empirical prediction of ground motion.
Bulletin of the Seismological Society of America, S43-S60.
Boore, D. M., Gibbs, J. F., Joyner, W. B., Tinsley, J. C., & Ponti, D. J. (2003).
Estimated ground motion from the 1994 Northridge, California, earthquake
at the site of the interstate 10 and La Cienega Boulevard bridge collapse,
West Los Angeles, California. Bulletin of the Seismological Society of
America, 93(6), 2737-2751.
Boore, D. M., Joyner, W. B., & Fumal, T. E. (1997). Equations for estimating
horizontal response spectra and peak acceleration from Western North
American earthquakes: a summary of recent work. Seismological Research
Letters, 68(1), 128-153.
Bullen, K. E., & Bolt, B. A. (1985). An Introduction to the Theory of Seismology.
Cambridge, UK: Cambridge University Press.
Campbell, K. W., & Bozorgnia, Y. (2008). NGA ground motion model for the
geometric mean horizontal component of PGA, PGV, PGD and 5% damped
linear elastic response spectra for periods ranging from 0.01 to 10 s.
Earthquake Spectra, 24(1), 139-171.
Chiles, J.-P., & Delfiner, P. (2012). Geostatistics: modeling spatial uncertainty.
Hoboken, NJ: Wiley.

42

Chiou, B., & Youngs, R. (2008). An NGA model for the average horizontal
component of peak ground motion and response spectra. Earthquake Spectra,
24(1), 173-215.
Cornell, C. A. (1968, October). Engineering Seismic Risk Analysis. Bulletin of the
Seismological Society of America, 58(5), 1583-1606.
Field, E. H. (2005). Probabalistic seismic hazard analysis (PSHA): A primer.
http://www.relm.org/tutorial_materials/PSHA_Primer_v2.pdf.
Field, E. H., Dawson, T. E., Felzer, K. R., Frankel, A. D., Gupta, V., Jordan, T. H., . .
. Wills, C. J. (2009, August). Uniform California Earthquake Rupture
Foreast, Version 2 (UCERF 2). Bulletin of the Seismological Society of
America, 99(4), 2053-2107.
Field, E. H., Gupta, N., Gupta, V., Blanpied, M., Maechling, P., & Jordan, T. H.
(2005). Hazard Calculations for the WGCEP-2002 earthquake forecast using
OpenSHA and distributed objecct technologies. Seismological Research
Letters, 76(2), 161-167.
Field, E. H., Jordan, T. H., & Cornell, C. A. (2003). OpenSHA: A developing
community-modeling environment for seismic hazard analysis. Seismological
Research Letters, 74(4), 406-419.
Gilbert, G. K., Holmes, J. A., Humphrey, R. L., Sewell, J. S., & Soule, F. (1907). The
San Francisco Earthquake and Fire of April 18, 1906 andtheir Effects on
Structures and Structural Materials. Washington: Government Printing
Office.
Goda, K., & Atkinson, G. M. (2009). Probabalistic characterization of spatially
correlated response spectra for earthquakes in Japan. Bulletin of the
Seismological Society of America, 99(5), 3003-3020.
Goda, K., & Hong, H. P. (2008). Spatial correlation of peak ground motions and
response spectra. Bulletin of the Seismological Society of America, 98(1), 354365.
Housner, G. W., & Jennings, P. C. (1982). Earthquake Design Criteria. Oakland,
CA: Earthquake Engineering Research Institute.
Jayaram, N., & Baker, J. W. (2009). Correlation model for spatially distributed
ground-motion intensities. Earthquake Engineering and Structural
Dynamics, 38, 1687-1708.
Joyner, W. B., & Boore, D. M. (1988). Measurement, characterization, and
prediction of strong ground motion. Earthquake Engineering and Soil
Dynamics (pp. 43-102). Park City: American Society of Civil Engineers.
Joyner, W. B., & Boore, D. M. (1993, April). Methods for regression analysis of
strong-motion data. Bulletin of the Seismological Society of America, 83(2),
469-487.
Kanamori, H. (1977). The energy release in great earthquakes. Journal of
Geophysical Research, 82(20), 2981-2987.
Kanamori, H. (1978). Quantification of Earthquakes. Nature, 271, 411-414.

43

Park, J., Bazzurro, P., & Baker, J. W. (2007). Modeling spatial correlation of ground
motion Intensity Measures for regional seismic hazard and portfolio loss
estimation. Application of Statistics and Probability in Civil Engineering.
Porter, K. A., Field, E. H., & Milner, K. (2012). Trimming the UCERF2 hazard logic
tree. Seismological Research Letters, 83(5), 815-828.
Reitherman, R. K. (2012). Earthquakes and Engineers. Reston, VA: ASCE Press.
Rezaeian, S., & Kiureghian, A. D. (2010). Stochastic modeling and simulation of
ground motions for performance-based earthquake engineering. PEER Report
2010/02, Pacific Earthquake Engineering Research Center.
Trifunac, M. D., & Brady, A. G. (1975, February). On the correlation of seismic
intensity scales with the peaks of recorded strong ground motion. Bulletin of
the Seismological Society of America, 65(1), 139-162.
Wang, M., & Takada, T. (2005). Macrospatial correlation model of seismic ground
motions. Earthquake Spectra, 21(4), 1137-1156.
Working Group on California Earthquake Probabilities (WGCEP). (2007). The
Uniform California Earthquake Rupture Forecast, Version 2 (UCERF 2).
U.S. Geol. Surv. Open-File Report 2007-1437.

44

APPENDIX

PGA (g)

NE-SW
(perp. faults)

NW-SE
(II to faults)

Distance
from main
site (km)
ID
2
3
4
5
6
7
8
9
10

lat (deg)

0 37.81501944
7.2
37.753916
15.2
37.687946
28.1
37.591687
17.9
37.952433
39.9
38.118378
7.2
37.841769
3.9
37.830458
5.3
37.797054
10.2

37.779793

50% in 10

50% in 50

10% in 50

2% in 50

Prob. Of Exc.

long (deg)

20

100

500

2500

Return Period

-122.3589639
-122.331946
-122.294463
-122.208561
-122.465202
-122.603299
-122.284069
-122.319123
-122.415389

0.06 % Diff
0.0604
1%
0.0611
2%
0.0632
5%
0.0561
7%
0.0504
16%
0.0635
6%
0.0619
3%
0.0572
5%

0.2217 % Diff
0.2178
2%
0.2099
5%
0.2019
9%
0.2174
2%
0.1877
15%
0.2505
13%
0.2358
6%
0.2088
6%

0.4673 % Diff
0.4627
1%
0.4515
3%
0.4462
5%
0.4684
0%
0.447
4%
0.6491
39%
0.5406
16%
0.4802
3%

0.691 % Diff
0.687
1%
0.6792
2%
0.6821
1%
0.6935
0%
0.6678
3%
1.0693
55%
0.8389
21%
0.7326
6%

0.1952

0.5798

0.9239

-122.466225 0.0546

9%

12%

24%

34%

Correlation Investigation: Hazard Curve Calculations for San Francisco Bay Area sites

PGA (g)

NE-SW
(II faults)

NW-SE
(perp. faults)

50% in 10

ID
1
2
3
4
5
6
7
8

Distance
from main
lat (deg)
long (deg)
20
site (km)
0 34.24109722 -118.5292722 0.071 % Diff
2.9
34.264421
-118.515997 0.0712
0%
1.6
34.254093
-118.521196 0.0712
0%
3.2
34.214772
-118.543261 0.0707
0%
12.9
34.134329
-118.583238 0.0685
4%
4.3
34.227779
-118.484874 0.0715
1%
11.6
34.211037
-118.407999 0.0722
2%
5.9
34.261235
-118.588275 0.0704
1%
2.8
34.252143
-118.556289 0.0708
0%

50% in 50

10% in 50

2% in 50

Prob. Of Exc.

100

500

2500

Return Period

0.2456 % Diff 0.6562 % Diff


0.2554
4% 0.7822
19%
0.2514
2% 0.7123
9%
0.2338
5% 0.5703
13%
0.2043
17% 0.4217
36%
0.2385
3% 0.6148
6%
0.2299
6% 0.5422
17%
0.2555
4% 0.6997
7%
0.2514
2% 0.6843
4%

1.0877 % Diff
1.3057
20%
1.1787
8%
0.9247
15%
0.6522
40%
1.2019
10%
0.8932
18%
1.1853
9%
1.1334
4%

Correlation Investigation: Hazard Curve Calculations for San Fernando Valley sites

PGA (g)

NE-SW
(perp.
faults)

NW-SE
(II to faults)

Distance
from main
site (km)
ID
1
2
3
4

lat (deg)

long (deg)

50% in 10

50% in 50

10% in 50

2% in 50

Prob. Of Exc.

20

100

500

2500

Return Period

0.0198 % Diff
0.0198
0%
0.0198
0%
0.0196
1%
0.0197
1%

0.1968 % Diff
0.1898
4%
0.2029
3%
0.1752
11%
0.2456
25%

0.7101 % Diff
0.6655
6%
0.7553
6%
0.5876
17%
1.1493
62%

0 35.142275 -90.0269111
0.06 % Diff
5.2 35.141917 -89.969209 0.0604 1108%
3.6 35.14751
-90.06648 0.0611 1122%
16.6 35.139248
-89.84451 0.0632 1164%
28.9 35.17129 -90.343395 0.0561 1022%

7.3 35.207683

-90.020792 0.0635

1170%

0.0207

5%

0.2119

8%

0.7981

12%

7.9 35.071061

-90.031804 0.0619

1138%

0.0188

5%

0.1798

9%

0.6295

11%

Correlation Investigation: Hazard Curve Calculations for Memphis sites

You might also like