You are on page 1of 185

SEISMIC HAZARD ANALYSIS AND SEISMIC SLOPE STABILITY

EVALUATION USING DISCRETE FAULTS IN NORTHWESTERN


PAKISTAN






BY

BYUNGMIN KIM








DISSERTATION

Submitted in partial fulfillment of the requirements
for the degree of Doctor of Philosophy in Civil Engineering
in the Graduate College of the
University of Illinois at Urbana-Champaign, 2012



Urbana, Illinois


Doctoral Committee:

Professor Youssef M.A. Hashash, Chair, Director of Research
Associate Professor Scott M. Olson, Co-director of Research
Professor Gholamreza Mesri
Professor Amr S. Elnashai
Associate Professor Junho Song

ii

ABSTRACT
The M
w
7.6 earthquake that occurred on 8 October 2005 in Kashmir, Pakistan, resulted in
tremendous number of fatalities and injuries, and also triggered numerous landslides. Although
there are no reliable means to predict the timing of the earthquake, it is possible to reduce the
loss of life and damages associated with strong ground motions and landslides by designing and
mitigating structures based on proper seismic hazard and seismic slope stability analyses. This
study presents the methodology and results of seismic hazard and slope stability analyses in
northwestern Pakistan.
The first part of the thesis describes the methodology used to perform deterministic and
probabilistic seismic hazard analyses. The methodology of seismic hazard analysis includes
identification of seismic sources from 32 faults in NW Pakistan, characterization of recurrence
models for the faults based on both historical and instrumented seismicity in addition to geologic
evidence, and selection of four plate boundary attenuation relations from the Next Generation
Attenuation of Ground Motions Project. Peak ground accelerations for Kaghan and
Muzaffarabad which are surrounded by major faults were predicted to be approximately 3 to 4
times greater than estimates by previous studies using diffuse areal source zones. Seismic hazard
maps for PGA and spectral accelerations at periods of 0.2 sec and 1.0 sec corresponding to 475-,
975-, and 2475-year return periods were produced for NW Pakistan. Based on deaggregation
results, a discussion of the conditional mean spectra for engineering applications is presented.
The second part of the thesis proposes factors that affect distribution of shallow
landslides triggered by an earthquake. Landslides are the most common consequence of
earthquakes, resulting in significant amount of damages of structures and lives. Significant
damage was induced from the landslides triggered by the 2005, Kashmir, Pakistan, earthquake.
Therefore, predicting locations and severity of landslide is an essential part of earthquake
engineering. However, the currently used seismic slope stability analysis cannot capture the
actual trend of landslide distribution, especially high landslide concentration near field. This
study proposes the effect of vertical ground acceleration, topographic effects, and bond break
effects, in addition to the strong horizontal ground acceleration, as factors that contribute to the
landslide distribution near earthquake source. Landslide database from four earthquake cases
(1989 Loma-Prieta, U.S.; 1999 Chi-Chi, Taiwan; 2005 Kashmir, Pakistan; and 2008 Wenchuan,
China) were selected to verify these factors for slope stability analysis.

iii












To my parents and grandparents














iv

ACKNOWLEDGEMENT
I would like to express my deep appreciation and respect to my advisor, Professor
Youssef M.A. Hashash, and my co-advisor, Professor Scott M. Olson, for their careful attention,
patient guidance, enthusiastic advice, and constant support throughout the period of study.
Having a chance to study under their supervision was one of the most fortunate things in my life.
I would like to extend my sincere appreciation to the members of the committee, Professors
Gholamreza Mesri, Amr S. Elnashai, and Junho Song for their insightful advice which
significantly enhanced this dissertation.
I also appreciate Professors Timothy D. Stark and James H. Long for their skillful and
passionate lectures that inspired me greatly. I deeply appreciate Professor Duhee Park at the
Hanyang University who led me to Illinois inspiring to aspire to higher education.
I would like to thank my gentle officemates, David Groholski, Camilo Phillips, Sungwoo
Moon, Michael Musgrove, for making a vibrant research environments with pleasant casual
chats as well as helpful discussion. Special thanks are extended to GESO members: Randa
Asmar, Fangzhou Dai, Joseph Harmon, Maria Ines Romero, Mark Muszynski, Janson Funk,
Andrew Tangsombatvisit, Mohamad Jammoul. I believe we will be all connected as we move
forward careers in our own professions, Oscar Moreno-Torres.
I am grateful to all members of the Korean Student Association in Civil Engineering,
especially to Seung Jae Lee, Junhan Kim, Robin Kim, Hongki Joe, Moochul Shin, Hyungchul
Yoon, Seungmin Lim, Junho Chun. I would also like to thank Professors Kyungsoo Park,
Sunghan Sim, Taekeun Oh.
Most importantly, I would like to thank my parents and younger brother for their love,
support, and encouragement. I specially thank my grandparents who passed away during the
period of my study. No words can express how grateful I am for their endless love and support
on me.
The supports from USAID and Higher Education Commission of Pakistan under Award
No. AID NAS PGA-7251 are gratefully acknowledged.

v

TABLE OF CONTENTS
LIST OF SYMBOLS ..................................................................................................................... vi
LIST OF ABBREVIATIONS ...................................................................................................... viii
Chapter 1. INTRODUCTION ..................................................................................................... 1
Chapter 2. PRIOR SEISMIC HAZARD STUDIES IN NW PAKISTAN .................................. 3
Chapter 3. DETERMINISTIC SEISMIC HAZARD ANALYSIS .............................................. 8
Chapter 4. PROBABILISTIC SEISMIC HAZARD ANALYSIS ............................................. 32
Chapter 5. SEISMIC HAZARD RESULTS .............................................................................. 54
Chapter 6. CURRENT SEISMIC SLOPE STABILITY APPROACH ..................................... 92
Chapter 7. FACTORS THAT AFFECT PSEUDO-STATIC SLOPE STABILITY ............... 111
Chapter 8. PROPOSED NEW APPROACH TO PSEUDO-STATIC SLOPE STABILITY
ANALYSIS AND VERIFICATION USING EARTHQUAKE CASES .............. 121
Chapter 9. CONCLUSIONS AND RECOMMENDATION FOR FUTURE RESEARCH ... 154
Appendix A. ESTIMATION OF MOHR-COULOMB STRENGTH PARAMETERS ............. 159
REFERENCES ........................................................................................................................... 168



vi

LIST OF SYMBOLS
a
h
horizontal acceleration
a
h,max
maximum horizontal acceleration
a
max
maximum acceleration
a
v
vertical acceleration
a
v,max
maximum vertical acceleration
a
y
yield acceleration
b slope of exponential recurrence model
b
l
slip rate at greater depths
b
s
slip rate at sesimogenic depths
c cohesion intercept of the failure mass
D depth to the bottom of rupture
d average displacement over the slip surface
D
t
failure mass thickness
F
h
horizontal pseudo-static force
F
v
vertical pseudo-static force
h slope height
i slope angle
k
bond
pseudo-static coefficient for bond break effects
k
h
horizontal pseudo-static coefficient
k
v
vertical pseudo-static coefficient
M earthquake magnitude
m
b
body wave magnitude
m
i
material constant for intact rock
M
L
local magnitude
M
s
surface wave magnitude
m
u
upper bound magnitude for the exponential recurrence model
M
w
moment magnitude
M
0
seismic moment
m
0

threshold magnitude for the exponential recurrence model
(m
c
) Characteristic rate

vii

R source-to-site distance
RA rupture area
R-factor seismogenic scaling factor
S average seismic slip rate
Sa Spectral acceleration
T spectral period
T
s
recurrence interval
T* period of interest
V
s,30
shear wave velocity in the upper 30 m
W weight of the failure mass
Z
tor
depth to the top of rupture
Z
1.0
depth to shear wave velocity of 1 km/s
Z
2.5
depth to shear wave velocity of 2.5 km/s
unit weight of the failure mass
ground motion uncertainty

m
annual activity rate of earthquakes of magnitude greater than m
rigidity of shear modulus which is usually taken to be 310
11
dyne/cm
2
activity rate, complementary cumulative earthquake rate for m > m
0

correlation coefficient of CMS
o
ci
uniaxial compressive strength of the intact rock
o
t
tensile strength
1
o'
major effective principal stress
3
o' minor effective principal stress
| friction angle of the failure mass


viii

LIST OF ABBREVIATIONS
AFPS French association for earthquake engineering
BAAS British association for the advancement of science
CMS conditional mean spectra
COV coefficient of variation
DF disturbance factor
DSHA deterministic seismic hazard analysis
EC European seismic code provision
FS factor of safety
GMPE ground-motion prediction equation
GSI geological strength index
ISC International seismological centre
KPSZ Kohat-Potwar-Salt range seismic zone
LC landslide concentration
LS least square method
MBT main boundary thrust
MCT main central thrust
MKT main karakoram thrust
ML maximum likelihood method
MMT main mantle thrust
NEHRP national earthquake hazards reduction program
NGA next generation of ground-motion attenuation models
NOAA national oceanic and atmospheric administration
PGA peak ground acceleration
PGV peak ground velocity
PMD Pakistan meteorological department
PHSZ Peshawar-Hazara seismic zone
PSHA probabilistic seismic hazard analysis
SA spectral acceleration
SASZ Swat-Astor seismic zone
SKSZ Surghar-Kurram seismiczone

ix

UHS uniform hazard spectra
USGS U.S. geological survey
V/H ratio vertical-to-horizontal acceleration ratio
WGCEP working group on California earthquake probabilities

1

CHAPTER 1. INTRODUCTION
1.1 Statement of the problem
We have been experiencing numerous catastrophic earthquakes all around the world. As
there are no reliable means to predict the timing of these earthquakes, engineers focus on
preventing significant earthquake loss by designing earthquake-resistant new structures and
seismically mitigating existing structures based on proper seismic hazard analysis. However
many of current seismic studies employ areal source zones for the region where no sufficient
information of faults is available [e.g. Africa (Mavonga and Durrheim 2009); Caribbean Islands
(Bozzoni et al. 2011); India (Jaiswal and Sinha 2007); Indonesia and Malaysia (Petersen et al.
2004); Portugal (Vilanova and Fonseca 2007); Spain (Mezcua et al. 2011); United Arab Emirates
(Aldama-Bustos et al. 2009)]. The prior seismic hazard studies for NW Pakistan were also
conducted based on the diffuse areal source zones, resulting in the distribution of seismic hazard
to known faults throughout the entire source area, and thus underestimating ground acceleration.
This may cause a serious problem to seismic design for socio-economically important cities
located close to major faults like Islamabad and Muzaffarabad.
Furthermore, the currently-used seismic slope stability analysis cannot explain the trend
of earthquake-induced landslides distribution which is an important aspect that needs to be
considered for seismic design. Specifically, the pseudo-static slope stability analysis approach
predicts high factor of safety near a fault, which is inconsistent with the observation of high
concentration of landslides near a fault. Therefore, it is necessary to investigate the factors that
affect the seismic slope stability, in addition to the strong horizontal acceleration.

1.2 Objectives and scope of study
The main purpose of this study is to present the methodology and results of seismic
hazard analysis using discrete faults and seismic slope stability analysis in NW Pakistan. To
achieve this goal, the deterministic and probabilistic seismic hazard analyses are performed for
selected major cities using input parameters developed based on available information on
seismicity and characteristics of discrete faults. The seismic hazard map corresponding to 475-,
975-, and 2475-year return periods are generated for the entire region of NW Pakistan. Use of the
conditional mean spectrum instead of the uniform hazard spectrum is also introduced.

2

To evaluate the regional performance of earthquake-induce landslide during the 2005
Kahsmir, Pakistan, earthquake, three possible factors are introduced: (1) vertical ground
accelerations; (2) topographic effects; and (3) bond break effects. The new approach using
these three factors is verified by landslide data from three additional earthquakes: 1989 Loma-
Prieta, U.S.; 1999 Chi-Chi, Taiwan; and 2008 Wenchuan, China earthquakes.
1.3 Organization of the thesis
Chapter 2 introduces prior seismic hazard studies in NW Pakistan, and describes the
disadvantage of these studies due to the use of diffuse areal source zone. Chapter 3 describes the
methodology of a deterministic seismic hazard analysis for NW Pakistan using discrete faults.
This chapter presents source characterization and selection of ground motion prediction models
as well as the result of sensitivity analysis and verification of methodology by comparing with
measurements from the 2005 Kahsmir, Pakistan, earthquake. Chapter 4 describes
characterization of recurrence models for a probabilistic seismic hazard analysis. The results
from both deterministic and probabilistic seismic hazard analyses are summarized in Chapter 5,
comparing with the results from prior studies using areal source zones. Based on deaggregation
results, a discussion of the conditional mean spectra for engineering applications is presented in
this chapter.
Chapter 6 introduces the characteristics of earthquake-induced landslides and genera
approaches to analyze seismic slope stability. This chapter also raises the problem of currently-
used pseudo-static slope stability analysis using landslide database from the 2005 Kahsmir,
Pakistan, earthquake. Chapter 7 proposes the possible factors that affect the slope stability,
resulting in better understanding of regional distribution of earthquake-induced landslides.
Chapter 8 verifies the new approach to pseudo-static slope stability analysis using four
earthquake cases.
Finally, conclusions are reached in Chapter 9, and possible future developments are
suggested.

3

CHAPTER 2. PRIOR SEISMIC HAZARD STUDIES IN NW PAKISTAN
2.1 Introduction
Several seismic hazard studies recently have been conducted for regions of Pakistan
because of its active tectonic setting, including studies by Monalisa et al. (2007), NORSAR and
Pakistan Meteorological Department (PMD) (2006), and PMD and NORSAR (2006). While
these studies incorporate historical and instrumented earthquakes, each involves some potentially
limiting assumptions regarding seismic source zone characterization and earthquake catalog
construction. This chapter summarizes the key features and drawbacks of these prior studies.
2.2 Monalisa et al. (2007)
Monalisa et al. (2007) presented a probabilistic seismic hazard analysis (PSHA) for the
NW Himalayan belt. Their earthquake catalog includes 1057 instrumentally-recorded
earthquakes (M
w
4) obtained chiefly from the International Seismological Centre (ISC) and
U.S. Geological Survey (USGS). Based on this and other geological information, Monalisa et al.
(2007) defined four diffuse seismic source zones: the Peshawar-Hazara, Surghar-Kurram, Kohat-
Potwar-Salt Range, and Swat-Astor seismic zones (PHSZ, SKSZ, KPSZ, and SASZ,
respectively). The corresponding recurrence relations, characterized by b-values, annual activity
rates, and upper bound magnitudes (m
u
), were developed for each seismic source zone. The
activity rate and b-value were estimated by earthquake data in each zone, and the upper bound
magnitude was estimated by the relationship between magnitude and rupture length (Bonilla et al.
1984; Nowroozi 1985; Slemmons et al. 1989; Wells and Coppersmith 1994). The depth for each
seismic zone was defined as the average focal depth of earthquakes in that zone.
Monalisa et al. (2007) performed a PSHA using the Ambraseys et al. (1996) and Boore et
al. (1997) attenuation relations to predict peak ground accelerations (PGA) for 10 cities,
including Islamabad and Muzaffarabad. Careful review of this study called attention to the
unexpectedly low ground motions. For example, the PGA with 475-year return period [i.e., 10%
probability of exceedance (PE) in 50 years] for Islamabad is only 0.10g and 0.15 g with the
attenuation relations of Ambraseys and Simpson (1996) and Boore et al. (1997), respectively,
despite Islamabad being located within 4 km of the seismically-active MBT. Similarly, the PGA
with 475-year return period is only 0.12g for Muzaffarabad, despite its proximity to the active

4

Jhelum Fault and Riasi Thrust. PGA values for other cities calculated by Monalisa et al. (2007)
are shown in Table 2.1.
The seismic hazard is likely underestimated because of the following reasons:
1. Diffuse areal source zones. Areal source zones are useful where faults are not well-
characterized. However, the source zones defined by Monalisa et al. (2007) are too diffuse to
capture seismicity adjacent to discrete faults. Furthermore, the source zones appear to have
neglected the east section of the MBT, which plays an important role in the regional seismic
hazard.
2. Incomplete earthquake catalog. The earthquake catalog is missing major events,
including the 1555 Kashmir, 1905 Kangra, and 2005 Kashmir earthquakes. In addition, the
instrumented earthquake catalog appears incomplete, as illustrated in Figure 2.1. Note that the
apparent increase in seismicity in both catalogs in the latter part of the 20
th
century results from
an increase in seismic monitoring, not an increase in actual seismicity.
3. Inconsistency between hazard curves and tabulated PGAs. Values of PGA read from
the hazard curves and those tabulated by Monalisa et al. (2007) did not match for many cities.
For example, the PGAs reported for Peshawar were 0.14g and 0.15g with 475-year return period
using the two attenuation relationships. However, the corresponding PGAs obtained from the
hazard curves were approximately 0.32g. Similarly, PGAs reported for Muzaffarabad were 0.10g
and 0.13g with 475-year return period, but the hazard curves showed PGAs of about 0.14g and
0.16g.
2.3 PMD and NORSAR (2006)
The Pakistan Meteorological Department (PMD) and NORSAR (2006) computed seismic
hazard for Azad Kashmir and northern Pakistan using a PSHA based on diffuse areal source
zones. The region was divided into 16 diffuse source zones based on tectonic setting and local
seismicity. The earthquake catalog used in their study was based chiefly on the NORSAR
database which was collected from agencies worldwide (M
w
4.5) (PMD and NORSAR 2006).
Another main source was the PMD database containing historical earthquakes from 1905 and
most recent instrumented earthquakes. Recurrence relations and focal depths for each source
zone were developed from earthquakes assigned to each zone. The attenuation model proposed
by Ambraseys et al. (1996) was used to establish the seismic hazard contour map of PGA values
for return periods of 100, 500, and 1000 years.

5

The PSHA results yielded low PGA values for Islamabad and Muzaffarabad (see Table
2.1). The PGAs for Islamabad for return periods of 500 years and 1000 years are 0.20g and
0.26g, respectively. The PGAs for Muzaffarabad are 0.20g and 0.31g for return periods of 500
and 1000 years, respectively. The likely reason for the low PGAs is that using diffuse areal
source zones tends to smear local seismicity corresponding to known faults throughout the
entire source area, thereby decreasing seismic hazard at locations proximate to active faults.
2.4 NORSAR and PMD (2006)
NORSAR and PMD (2006) performed PSHA for the cities of Islamabad and Rawalpindi.
The earthquake data obtained from the databases of the ISC, USGS, EHB, NORSAR, and PMD
were used. Based on seismo-tectonic setting, 11 shallow area source zones that cover shallow
earthquakes, and one deep zone that covers deep earthquakes in Hindu Kush region were defined.
However the deep source zone was ignored in their study because it is far from the target cities
and does not affect the hazard results. Like other studies (PMD and NORSAR 2006; Monalisa et
al. 2007), recurrence relations were developed for each source zone using the earthquakes
assigned to each zone. In addition to area source zones, NORSAR and PMD (2006) modeled the
Jhelum fault (a strike-slip fault in NW Pakistan) in two segments, and they used the Ambraseys
et al. (1996) attenuation relation.
PGA values were reported for a grid of 20 points in a single seismic zone, covering
Islamabad and Rawalpindi, and values ranged from 0.19g to 0.21g with 500-year return period.
These PGAs again are quite low considering that the seismically-active MBT crosses the study
area.
2.5 Summary
The prior seismic studies for NW Pakistan used diffuse areal source zones, resulting in
the PGAs of about 0.1 to 0.2g for 475- or 500-year return period for Muzaffarabad. Considering
that Muzaffarabad is the city that was severely affected by the 2005 Kashmir earthquake and
located proximate to active faults, this low PGA prediction for Muzaffarabad is unreasonable.
PGAs estimated for other cities also rarely exceed 0.2g. Another observation is that PGAs show
little variation among cities regardless of their proximity to faults.


6

Table 2.1 PGAs predicted by Monalisa et al. (2007) for 475-year return period using attenuation
relations by Ambraseys et al. (1996) and Boore et al. (1997), and by PMD and NORSAR (2006)
for 500-year return period using attenuation relation by Ambraseys et al. (1996).
Site
PGA (g)
Monalisa et al. (2007) PMD and NORSAR (2006)
Ambraseys et al. (1996) Boore et al. (1997) Ambraseys et al. (1996)
Astor 0.07 0.082 0.28
Bannu 0.06 0.08 0.08
Islamabad 0.10 0.15 0.20*
Kaghan 0.09 0.12 0.20
Kohat 0.20 0.21 0.13
Malakand 0.20 0.21 0.20
Mangla 0.16 0.18 0.10
Muzaffarabad 0.14

0.16

0.20
Peshawar 0.32

0.32

0.20
Talagang 0.15 0.16 0.12
*NORSAR and PMD (2006) reported PGA values for Islamabad ranging from 0.19g to 0.21g
with 500-year return period.

Values from the hazard curves, which are inconsistent with tabulated values in Monalisa et al.
(2007).


7


Figure 2.1 Instrumented earthquake (M
w
4.0) from the composite catalog used in this study
with time compared with catalog used by Monalisa et al. (2007). The area considered is 69.5 E
75.5 E, 31.5 N 36.5 N.


Time
(year)
0
200
400
600
800
N
u
m
b
e
r

o
f

e
v
e
n
t
s
<
1
9
1
2
1
9
1
2
-
1
9
2
1
1
9
2
2
-
1
9
3
1
1
9
3
2
-
1
9
4
1
1
9
4
2
-
1
9
5
1
1
9
5
2
-
1
9
6
1
1
9
6
2
-
1
9
7
1
1
9
7
2
-
1
9
8
1
1
9
8
2
-
1
9
9
1
1
9
9
2
-
2
0
0
2
>
2
0
0
2
Monalisa et al. (2007)
This study

8

CHAPTER 3. DETERMINISTIC SEISMIC HAZARD ANALYSIS
3.1 Introduction
DSHA has an advantage in that less information is required for fault recurrence rates
which is not widely available for NW Pakistan. However, DSHA cannot account for the
probability of occurrence of earthquakes.
The four steps for a typical deterministic seismic hazard analysis (Reiter 1990) are
illustrated in Figure 3.1 and include:
1) Characterization of all possible earthquake zones. This step includes definition of
geometry and maximum potential magnitudes of source zones.
2) Selection of source-to-site distance. Generally, the shortest distance to rupture or the
shortest horizontal distance from source to site is used as a distance, depending on
attenuation relationships also known as ground-motion prediction equations (GMPEs).
3) Selection of the controlling earthquake source in terms of the ground motion
parameters. In this step, GMPEs are used the ground motion parameters such as
acceleration and velocity at a given magnitude and distance from the source to the site.
Comparing all possible source zones, one controlling source zone is selected.
4) Computation of seismic ground motions parameters at the site. Usually, ground
acceleration is used.

This chapter describes the methodology of a deterministic seismic hazard analysis
(DSHA) (Reiter 1990) for NW Pakistan using information available for individual, discrete
faults, rather than diffuse source zones. The tectonic settings and the earthquake catalog
compilation were introduced, followed by explanation on fault characterization and selection of
GMPEs. This chapter also presents the results of sensitivity analyses for input parameters and
validates the developed methodology by comparison with measurements from the 2005 Kashmir,
Pakistan, earthquake.
3.2 Tectonic settings
Pakistan is located in a highly seismic area and has experienced large and destructive
earthquakes historically and in recent times. The most recent severe earthquake occurred on 8
October 2005 in Kashmir, with a moment magnitude of 7.6 and focal depth of 26 km, and was

9

followed by numerous aftershocks (U.S. Geological Survey (USGS) 2010). This earthquake
resulted in at least 86,000 fatalities and more than 69,000 injuries (USGS 2010). The earthquake
also triggered over 2,400 landslides, extensive liquefaction in alluvial valleys, and numerous
building collapses (Bhat et al. 2005; Jayangondaperumal et al. 2008; Aydan et al. 2009). These
consequences in a moderately populated region brought renewed attention to seismic hazards in
Pakistan.
Pakistan lies on the Indian Plate, which forms the Himalayan mountain ranges and
flexures due to collision into the Eurasian Plate at a rate of about 45 mm/year and rotates
counterclockwise (Sella et al. 2002) (Figure 3.2). This rotation and translation of the Indian Plate
causes left-lateral slip in Baluchistan at 42 mm/year and right-lateral slip in the Indo-Burman
ranges at 55 mm/year (Bilham 2004). The two main active fold-and-thrust belts in northwestern
Pakistan region are the Sulaiman belt and the northwest (NW) Himalayan belt as shown in
Figure 3.2. The Sulaiman mountain belt at the northwestern margin of the Indian subcontinent
contains the Chaman fault where the 2008 Balochistan earthquake (M
w
6.4) occurred (Monalisa
and Qasim Jan 2010). This focuses on the NW Himalayan belt (area highlighted in Figure 3.2).
Faults along the Himalayan belt have produced countless earthquakes, including at least
four great (M
w
> 8) earthquakes in 1505, 1803, 1934, and 1950. Bilham and Ambraseys (2005)
suggested that there is missing slip in the Himalayas, reporting that the calculated slip rate of
less than 5 mm/yr from the earthquake data over the past 500 years is less than one-third of the
observed slip rate (18 mm/yr) in the past decade. This missing slip is equivalent to four about M
w

8.5 earthquakes that could occur in the near future (Bilham and Ambraseys 2005). The NW
Himalayan belt includes several major thrusts, such as the MMT (Main Mantle Thrust) and MBT
(Main Boundary Thrust), which have produced several great earthquakes. The faults and
seismicity in NW Himalayan belt are shown in Figure 3.3.

3.3 Earthquake catalog
For this study, a new earthquake catalog has been complied, incorporating historical and
instrumented seismicity. Historical earthquakes were obtained from the earthquake database of
the National Oceanic and Atmospheric Administration (NOAA) and reviewed based on the
literature. Table 3.1 provides details related to the historical earthquakes incorporated in the
current catalog.

10

Figure 2.1 includes instrumented earthquakes in NW Pakistan from catalogs available
from the USGS, British Association for the Advancement of Science (BAAS), International
Seismological Summary (ISS; 1918~1963), International Seismological Centre (ISC;
1964~present), and PMD. These catalogs use different magnitude scales such as M
w
, surface
wave magnitude (M
s
), body wave magnitude (m
b
), and local magnitude (M
L
). These magnitudes
were converted to M
w
using conversions from Grunthal and Wahlstrom (2003) for M
L
and M
s
,
and conversions from Johnston (1996) and Hanks and Kanamori (1979) for m
b
. Table 3.2
summarized the conversions used in this dissertation. Figure 2.1 compares the catalog used in
this dissertation to that used by Monalisa et al. (2007). The discrepancy in these two catalogs
results from using different source information. Table 3.3 summarizes the M
w
> 6 instrumented
earthquakes in NW Pakistan. Duplicate and manmade events among these catalogs were
removed while developing a combined historical and instrumented catalog (1505~2006).
3.4 Source characterization
Based on the literature and regional tectonic settings, 32 faults in NW Pakistan were
identified as shown in Figure 3.3. The earthquakes in northwestern corner of Figure 3.3 were
assigned to the Main Karakoram Thrust (MKT) which is not included in this study because this
fault is far from the major cities in NW Pakistan, and therefore, will not significantly affect the
seismic hazard for these cities. Because the identified faults in NW Pakistan are densely located,
the surface projections of these faults cover most of earthquakes illustrated in the figure. Some
earthquakes that were difficult to attribute to a specific fault were assigned to the nearest fault,
thus a background source model was not considered in addition to the fault models.
3.4.1 Fault geometry
Table 3.4 summarizes the geometries for the 32 fault segments considered in this study.
Most of characterized faults in NW Pakistan are thrust faults created by compressional stresses
due to plate convergence. Other faults, including the Darband, Hissartand, Kund and Nowshera
faults, were considered to be normal faults based on the difference of plate motion velocities (52
mm/year and 38 mm/year) measured by Bendick et al. (2007) as shown in Figure 3.3. The
USGS divides large faults in California into multiple segments to account for differences in slip
rate, maximum magnitude, fault geometry, etc. On this basis, the MBT, one of the chief faults in

11

the region, was divided into a west segment striking E-W and an east segment striking NW-SE.
These segments are subject to different tectonic mechanisms due to their differing orientations.
The lengths of faults range from 35 km to 340 km. The other parameters needed to define
the fault geometry are dip, depth to top of rupture (Z
tor
), and depth to bottom of rupture (D), and
are illustrated in Figure 3.4. Faults striking from west to east were assumed to dip northward,
while faults striking northwest to southeast were assumed to dip northeastward due to the
movement of Indian Plate which subducts underneath the Eurasian Plate (McDougall and Khan
1990; Mukhopadhyay and Mishra 2005; Bendick et al. 2007; Raghukanth 2008). It was assumed
that thrust faults dip at 30, while normal and reverse faults dip at 60 due to lack of information
on fault dip angle. The Dijabba, Jhelum, and Kalabagh faults are strike-slip (McDougall and
Khan 1990; Dasgupta et al. 2000) and were assumed to dip at 90. Since major cities are located
on the footwall side of faults, the influence of dip angles on seismic hazard for these cities is
limited, although further study of dip angle may improve the seismic hazard characterization for
hanging wall zones.
Many of the faults in NW Pakistan lack sufficient characterization to assign other
geometric parameters such as top of rupture depth, Z
tor
, and bottom of rupture depth, D.
Therefore, the geometry of well-studied faults in the western U.S. was considered as analogs for
some faults in NW Pakistan. Many faults in NW Pakistan and the western U.S. are shallow
crustal faults located along plate boundaries and are of similar age (McDougall and Khan 1990;
Searle 1996; Dipietro et al. 2000). The northwestern U.S. is located in a collision zone forming
the Cascadia subduction zone, resulting in numerous shallow crustal reverse faults. While the
San Andreas fault is largely a transform fault, there are several compressional regimes that
produce oblique and reverse B-type faults. (B-type faults are major faults with measurable slip
rate but inadequate information on segmentation, displacement or date of last earthquake)
(WGCEP 2008). While this comparison is limited, it is anticipated that these faults are
reasonably analogous to provide missing information on Z
tor
and D. Based on this assumption,
yet limited, analog, identical values of Z
tor
and logic tree branch weights were assigned to all 32
fault segments. The logic tree branch weights assigned here were used by Petersen (2008) for the
USGS seismic hazard mapping project. Specifically, Z
tor
= 0, 2 km, and 4 km were assigned
based on M
max
for each fault. The logic tree branch weights will be discussed in Section 3.6. A
variable depth D was used, with values of 10 km, 15 km, and 20 km, based on the range of D

12

used by USGS databases for Intermountain West, Pacific Northwest, and California faults
(Petersen 2008) and the fault models of Himalayan Frontal Thrust, MBT, and MCT (Main
Central Thrust) in the region of India (Seeber and Armbruster 1981).
3.4.2 Maximum Magnitude, M
max

Tocher (1958) first suggested a correlation between measured earthquake magnitude and
rupture parameters such as length and displacement. More recently, Wells and Coppersmith
(1994) collected source parameters for 421 historical earthquakes and developed empirical
relationships between measured magnitude and rupture area (RA), rupture length, and rupture
width. These correlations are often used to estimate the maximum magnitude (M
max
) that a fault
is capable of producing. Hanks and Bakun (2002; 2008) and the Working Group on California
Earthquake Probabilities (WGCEP) (2003) suggested that the Wells and Coppersmith (1994)
correlation slightly underestimates M
max
for large RAs and proposed alternate relationships. In
this study, M
max
was estimated using both relationships proposed by and Hanks and Bakun (2008)
and the Ellsworth B correlation (WGCEP 2003), identical to the approach used by the USGS
(Petersen 2008) and WGCEP (2008). The M
max
values were then averaged, and this averaged
M
max
was used directly in the DSHA, and was assigned as the upper bound magnitude, m
u
, in the
PSHA.
3.4.3 R-factor
WGCEP (2003) used a seismogenic factor, R, to account for aseismic slip, i.e., fault creep
not contributing to seismicity. The R-factor is calculated as:

l s
b b factor R = 1 Eq. (3-1)

where b
s
is slip rate at seismogenic depths and b
l
is slip rate at greater depths. Values of b
s
can be
estimated from geodetic data and b
l
can be estimated from either geological or geodetic data. An
R-factor of unity implies that all fault slip occurs as earthquakes, while R-factor = 0 implies that
all fault slip occurs as aseismic creep. Multiplying the computed fault RA by R-factor reduces
the effective rupture area, and in turn, decreases the effective M
max
. Singh (2000) suggested that
there is an aseismic zone in the region lying between 35N-40N and 66E-76E which is
adjacent to the current study area. Therefore, it is possible for faults in NW Pakistan to

13

experience aseismic creep. Due to lack of geologic and geodetic information for faults in NW
Pakistan, R-factors of 0.8, 0.9, and 1.0 were assigned to the NW Pakistan faults with logic tree
branch weights in this study based on the values assigned to California faults (WGCEP 2003;
WGCEP 2008).
3.4.4 Comparison with historical earthquake records
Table 3.4 provides M
max
values computed from M
max
-RA relationships (where RA is
adjusted by R-factors). These values agree well with historical earthquakes. Most of the faults
have M
max
values ranging from 7.1 to 7.9, comparable to the major historical earthquakes in the
catalog (Table 3.1; M
w
~ 6.3 to 7.8). Faults with computed M
max
> 8, i.e., MBT (west), MBT
(east), MMT, and MCT, are located along the Himalayan belt, and these M
max
values are
consistent with the earthquake magnitudes (M
w
~ 8.5) suggested by Bilham and Ambraseys
(2005).
3.5 Attenuation relationships and parameters
Since no region-specific ground motion prediction model is available for Pakistan, The
five ground-motion prediction equations (GMPEs) developed in the Next Generation of Ground-
Motion Attenuation Models (NGA) project (Power et al. 2008) were considered. These GMPEs
are the Abrahamson-Silva (2008), Boore-Atkinson (2008), Campbell-Bozorgnia (2008), Chiou-
Youngs (2008), and Idriss relations (2008), all of which provide horizontal PGA, peak ground
velocity (PGV), and 5% damped elastic pseudo-response spectral accelerations in the period
range of 0 to 10 seconds for shallow crustal earthquakes. These GMPEs were developed using
significant earthquakes along plate boundaries including the 1995 Kobe, Japan and 1999 Chi-
Chi, Taiwan earthquakes as well as numerous earthquakes in the western U.S. As discussed
earlier, NW Pakistan is also located in a plate convergence region, and contains many shallow
crustal faults including the MBT and MMT. Therefore, applying the NGA GMPEs for NW
Pakistan was reasonable.
For all of the GMPEs, a Site Class B/C boundary condition in the upper 30m (shear wave
velocity in the upper 30m, V
s,30
= 760 m/s) was assumed. The Abrahamson and Silva (2008) and
Chiou and Youngs (2008) GMPEs require a depth where V
s
= 1 km/s, termed Z
1.0
. For V
s,30
=
760 m/, Z
1.0
= 32 m and 24 m were computed, respectively, using their recommended

14

correlations. The Campbell and Bozorgnia (2007) relation requires a depth where V
s
= 2.5 km/s,
Z
2.5
. For V
s,30
= 760 m/s, Z
2.5
= 0.63 km was computed using their recommended correlation.
Figure 3.5 compares the PGAs predicted using V
s,30
= 760 m/s. As expected, the
predicted PGAs for the hanging wall side of a thrust fault exceed the foot wall values until the
source-to-site distance exceeds the distance to the surface of the fault plane. The Chiou and
Youngs (2008) relation yields significantly larger PGAs than the other attenuation relations at
distances to fault less than 15 km; therefore, their relation was not utilized in this study. In all
seismic hazard analyses performed in this study, ground motion parameters computed using the
remaining four GMPEs were equally weighted.
For this regional study, directivity effects were not incorporated. The NGA-West 2
project is currently underway to incorporate directivity in the NGA GMPEs, with the intent to
improve on the directivity formulation proposed by Somerville (1997). Future use of these new
GMPEs or the use of region-specific GMPEs will improve the seismic assessment in NW
Pakistan.
3.6 DSHA logic tree
In a DSHA, a logic tree is used to consider alternative options for input parameters and
predictive models used to estimate seismic hazard. Figure 3.6 presents the logic tree and branch
weights used for all DSHA performed in this study. Weights of 0.4, 0.4, and 0.2 were assigned to
D = 10, 15, and 20 km, respectively based on the data for Quaternary faults in the western U.S
(Petersen 2008) and the fault models of Himalayan Frontal Thrust, MBT, and MCT in the region
of India (Seeber and Armbruster 1981). Weights of 0.2, 0.6, and 0.2 were assigned to R-factor =
0.8, 0.9, and 1.0, respectively, based on data from California faults (WGCEP 2008). Lastly, the
four selected GMPEs were equally weighted.
3.7 Controlling faults at selected sites
For locations proximate to multiple individual faults (or fault segments), the DSHA was
repeated for each fault segment to define the maximum ground motion parameter predicted at the
subject location and to identify the corresponding controlling fault. Table 3.5 lists the controlling
faults that yield the highest amplitude spectral accelerations at all periods at selected cities in
NW Pakistan. Because of the similarities of most M
max
values, the controlling fault for most

15

cities considered here is the most proximate fault. However, for some cities like Astor and
Peshawar, more distant faults with larger M
max
values are the controlling faults.
3.8 DSHA sensitivity analysis
Sensitivity analyses were conducted to understand the influence of a number of
parameters on the DSHA results. The sensitivity analysis using the depth to the bottom of rupture
(D) of 10 km, 15 km, and 20 km revealed that D did not significantly influence the computed
hazard for Islamabad and Muzaffarabad. These cities are located on the footwall of the
controlling faults, and D does not influence the distance to fault for sites located on the footwall.
For sites such as Kaghan, located on a hanging wall, the distance to fault does not change if the
site is within the horizontal projection of controlling fault, and therefore, D again has little
influence on seismic hazard (e.g., median PGAs of 0.73g and 0.70g for Kaghan were obtained
using D=10 km and 20 km, respectively). A sensitivity analysis of thrust fault dip angle was also
conducted, considering dip angles of 20 and 40. This analysis revealed that dip angle (within
this range) also has no influence on seismic hazard for most of the major cities located on a
footwall. The dip angle has a minor influence on seismic hazard for cities located on a hanging
wall. The median PGAs of 0.74g and 0.70g were computed for dip angles of 20 and 40,
respectively for Kaghan. This is primarily because Kaghan is located very close to the fault, and
dip angle variation does not significantly affect the closest distance to the fault. On the other
hand, Peshawar is located on a hanging wall and approximately 40 km away from the fault. Even
though of dip angle variation will affect the closest distance to the fault, the difference in
computed PGAs is minor due to the small values of PGA. The median PGAs of 0.23g and 0.33g
were computed for dip angles of 20 and 40, respectively for this city. Similarly, the R-factor
has little influence on the seismic hazard, at least for the small range of R values applied in this
study. Varying R from 0.8 to 1.0 only increases M
max
by approximately 0.1 magnitude units. For
Islamabad, an increase in R from 0.8 to 1.0 increased the median PGA from 0.50g to 0.51g.
Figure 3.7 illustrates the DSHA sensitivity to the GMPEs. As illustrated in the figure, the
Idriss (2008) GMPE predicts higher PGAs while the other three GMPEs (Abrahamson and Silva
2008; Boore and Atkinson 2008; Campbell and Bozorgnia 2008) predict similar PGAs. As
expected, the weighted PGAs are close to the PGAs predicted by these latter three relations.

16

3.9 Comparison with measurements from the 2005 Kashmir earthquake
As an initial validation of the fault model developed here, The measured accelerations
during the 2005 Kashmir earthquake predicted by DSHA were compared with those reported by
Durrani et al. (2005). In order to predict the ground motion based on the methodology proposed
in this study, all parameters and weights proposed in this thesis were considered except for the
location of rupture. The 2005 Kashmir earthquake rupture was approximately 60 km long, with a
focal depth of about 26 km (USGS 2010). The focal depth was assumed as the depth to the
bottom of rupture because most of the aftershocks occurred at shallower depths (Bendick et al.
2007) and other studies also reported similar depths to the bottom of rupture (Durrani et al. 2005;
Bendick et al. 2007). Using a dip angle of 30 as assumed for thrust faults in NW Pakistan,
maximum magnitudes of 7.60, 7.66, and 7.71 were computed for R-factors of 0.8, 0.9, and 1.0,
respectively. These magnitudes are consistent with the reported magnitude (M
w
7.6); therefore,
these computed magnitudes in the DSHA were used with branch weights shown in Figure 5. Z
tor

= 0 was set because M
w
> 7 as described in Figure 3.6. Because site conditions at the recording
stations are not known, subsurface conditions at each recording station were assumed to
correspond to Site Class C (very dense soil and soft rock) with V
s,30
= 560 m/s and Site Class D
(stiff soil) with V
s,30
= 270 m/s. The DSHA was performed for both Site Classes, and for V
s,30
=
560 m/s and 270 m/s, Z
1.0
= 130 m and 500 m were computed, respectively (Abrahamson and
Silva 2008), and Z
2.5
= 0.99 km and 2.3 km, respectively (Campbell and Bozorgnia 2008).
Figure 3.8 presents contour maps of median and 84
th
percentile (median plus one
standard deviation, o) PGA, respectively, for Site Class B/C boundary conditions corresponding
to V
s,30
= 760 m/s. Figure 3.9 shows the predicted PGA attenuation with distance modified to
Site Class C (V
s,30
= 560 m/s) and Site Class D (V
s,30
= 270 m/s) conditions to facilitate
comparison to the measured values. The measured values include two PGAs (i.e., two orthogonal
directions) at recording stations in Abbottabad, Murree, and Nilore. These data are plotted with a
source-to-site distance uncertainty of 10 km. In addition, horizontal PGAs were recorded on
rock at Tarbela Dam and at the base of the Barotha Power Complex. The measurement in
Tarbela is suspected to correspond to Site Class B. The mean of all median predictions 1 for
V
s,30
= 560 m/s envelopes the recordings except for the Nilore station. The Nilore measurement
was likely affected by the response of the raft foundation where the instrument was placed
(Durrani et al. 2005). Note that for V
s,30
= 270 m/s, the Idriss (2008) GMPE was not used

17

because it does not apply to V
s,30
< 450 m/s. In this case, the remaining three GMPEs were
equally weighted. Overall, it is considered that the comparison is acceptable.
3.10 Summary and discussion
In this chapter, the methodology of a deterministic seismic hazard analysis (DSHA)
(Reiter 1990) was described for NW Pakistan using information available for individual, discrete
faults, rather than diffuse source zones. 32 discrete faults were identified in NW Pakistan based
on the literature and regional tectonic settings. Many of these faults lack sufficient
characterization to assign geometric parameters such as Z
tor
, D, dip, R-factor which were
assigned based on assumption and analogs with western U.S. faults. Then the maximum
magnitudes for faults were estimated using the relationships between rupture area and maximum
magnitude, and were compared well with historical earthquakes and earthquake magnitudes
suggested by Bilham and Ambraseys (2005). Four GMPEs were selected from NGA project that
were developed using earthquakes along plate boundaries. To account for uncertainties, the logic
tree was used for input parameters and GMPEs. The sensitivity analyses revealed that the
geometric parameters such as Z
tor
, D, dip, R-factor do not significantly influence the computed
seismic hazard. The measured ground accelerations from the 2005 Kahsmir, Pakistan, earthquake
were compared with those predicted by DSHA. Because site conditions at the recording stations
are not known, the subsurface conditions at each recording station were assumed to correspond
to Site Classes C and D. The mean values of all median predictions 1 were in good
agreement with the recorded accelerations.


18

Table 3.1 Major historical earthquakes incorporated in the earthquake database for NW Pakistan.
Earthquake Year
Estimated
M
w

Estimated
Max.
MMI
Details Source
Paghman,
Afghanistan
(34.60N,
68.93E)
1505 7.0 - 7.8 IX - X
60-km rupture of
Chaman fault. Strike-
slip and dip-slip
movement
Quittmeyer and
Jacob (1979)
Kashmir,
Pakistan
(33.50N,
75.50E)
1555 7.6 n/a
Limited intensity
reports available
Ambraseys and
Douglas (2004)
Alingar Valley,
Afghanistan
(34.83N,
70.37E)
1842 n/a VIII - IX
Several hundred
fatalities in Alingar
River valley and
Jalalabad Basin
Quittmeyer and
Jacob (1979)
Kashmir,
Pakistan
(34.60N,
74.38E)
1885 6.3 VIII - IX
3,000 fatalities.
Numerous buildings
destroyed; large
fissures observed;
large landslide
triggered south of
Baramula
Lawrence (1967)
Quittmeyer and
Jacob (1979)
Bilham (2004)
Chaman,
Pakistan
(30.85N,
66.52E)
1892 6.75 IIV-IX
30-km rupture of
Chaman fault; Left-
lateral strike-slip
movement
Heuckroth and
Karim (1973)
Quittmeyer and
Jacob (1979)
Kangra, India
(33.00N,
76.00E)
1905 7.830.18 n/a
20,000 fatalities;
100,000 buildings
destroyed; limited
instrumented data
available
Ambraseys and
Bilham (2000)
Kaul (1911)




19

Table 3.2 Magnitude conversions used in this study.
Magnitude conversion relations
Grunthal and Wahlstrom (2003) ( ) ( ) ( )
2
013 . 0 046 . 0 08 . 0 56 . 0 11 . 0 67 . 0
L L w
M M M + + =
Grunthal and Wahlstrom (2003)
s w
M M =
Johnston (1996) ( )
2
0
077 . 0 679 . 0 28 . 18 log
b b
m m M + + =
Hanks and Kanamori (1979) ( ) 7 . 10 log 3 2
0
= M M
w


20

Table 3.3 Major instrumented earthquakes from the composite earthquake catalog used for this
study (M
w
6.0).
Year Latitude Longitude Magnitude (M
w
)
1914 32.80 N
75.30 E
6.2
1928 35.00 N
72.50 E
6.0
1972 36.00 N
73.33 E
6.4
1974 35.10 N
72.90 E
6.1
1981 35.68 N
73.60 E
6.3
1992 33.35 N
71.32 E
6.0
2002 35.34 N
74.59 E
6.4
2004 33.00 N
73.10 E
6.6
2005 34.63 N
73.63 E
7.6
2005 34.90 N
73.15 E
6.4
2005 34.75 N
73.20 E
6.2
2005 34.76 N
73.16 E
6.1



21

Table 3.4 Fault parameters for 32 fault segments considered in this study. Maximum magnitude computed based on Z
tor
= 0, D = 15
km, and R-factor = 0.9. R, N, and SS denote reverse, normal, and strike-slip, respectively.
Fault
ID
Fault name Type
Dip
()
Length
(km)
Max.
magnitude
(M
w
)
Slip rate
(mm/yr)
Activity rate
(year
-1
)
Characteristic rate
(year
-1
)
1 2 A B C
1 Balakot Shear Zone R 60 44 6.9 0.56 0.824 0.000 0.00048 0.00040 0.00049
2 Batal Thrust R 30 65 7.4 0.83 0.745 0.133 0.00068 0.00057 0.00036
3 Bhittani Thrust R 30 39 7.2 0.17 0.118 0.067 0.00017 0.00016 0.00011
4 Darband Fault N 60 51 7.0 1.21 0.157 0.000 0.00095 0.00008 0.00095
5 Dijabba Fault SS 90 83 7.2 0.37 0.549 0.600 0.00033 0.00122 0.00022
6 Himalayan Frontal Thrust R 30 187 8.0 2.39 0.078 0.067 0.00175 0.00014 0.00046
7 Hissartang Fault N 60 129 7.5 3.08 0.157 0.167 0.00221 0.00034 0.00119
8 Jhelum Fault SS 90 138 7.5 3.30 3.960 0.267 0.00234 0.00235 0.00135
9 Kalabagh Fault SS 90 53 7.0 0.24 0.078 0.033 0.00022 0.00009 0.00020
10 Karak Thrust R 60 80 7.3 0.36 0.157 0.067 0.00032 0.00018 0.00020
11 Khair-I-Murat Fault N 60 152 7.6 4.00 0.471 0.500 0.00279 0.00102 0.00137
12 Khairabad Fault R 60 225 7.8 2.00 0.118 0.100 0.00149 0.00022 0.00057
13 Khisor Thrust R 30 96 7.6 0.43 0.196 0.200 0.00038 0.00041 0.00014
14 Kotli Thrust R 30 65 7.4 0.83 0.039 0.033 0.00067 0.00007 0.00036
15 Kund Fault N 60 85 7.3 2.03 0.039 0.033 0.00151 0.00007 0.00108
16 Kurram Thrust R 30 148 7.8 0.66 0.235 0.267 0.00055 0.00054 0.00015
17 MCT R 30 333 8.3 4.25 1.451 0.600 0.00294 0.00166 0.00054
18 Mansehra Thrust R 30 53 7.3 0.68 2.118 0.134 0.00056 0.00125 0.00034
19 Marwat Thrust R 30 35 7.1 0.16 0.235 0.000 0.00015 0.00011 0.00011
20 MBT west R 30 225 8.1 4.00 0.392 0.400 0.00279 0.00083 0.00068
21 MBT east R 30 333 8.3 4.25 1.803 0.23 0.00294 0.00124 0.00054
22 MMT R 30 340 8.3 4.34 7.373 1.833 0.00300 0.00651 0.00054
23 Nathiagali Thrust R 30 59 7.4 0.27 0.196 0.033 0.00024 0.00015 0.00012
24 Nowshera Fault N 60 80 7.3 1.92 0.392 0.467 0.00144 0.00093 0.00106
25 Punjal Thrust R 30 103 7.7 2.46 0.627 0.500 0.00180 0.00110 0.00075
26 Puran Fault R 60 102 7.4 1.31 4.471 0.533 0.00102 0.00303 0.00060
27 Raikot Fault R 60 63 7.1 0.80 0.706 0.000 0.00065 0.00034 0.00053
28 Riasi Thrust R 30 105 7.7 1.34 1.569 0.300 0.00104 0.00124 0.00040
29 Riwat Thrust R 30 38 7.2 0.17 0.275 0.333 0.00016 0.00066 0.00011
30 Salt Range Thrust R 30 193 8.0 0.87 0.196 0.200 0.00070 0.00041 0.00016
31 Stak Fault R 60 62 7.1 0.79 0.824 1.167 0.00065 0.00226 0.00053
32 Surghar Range Thrust R 30 67 7.4 0.30 0.235 0.167 0.00027 0.00038 0.00013

22

Table 3.5 Controlling faults and closest source-to-site distance for DSHA conducted for major
cities in NW Pakistan.
Cities Fault number Fault name Closest distance (km)
Astor 22 MMT 23.91
Balakot 8 Jhelum Fault 1.94
Bannu 10 Karack Fault 8.57
Islamabad 20 MBT (west) 3.14
Kaghan 17 MCT 3.42
Kohat 20 MBT (west) 1.55
Malakand 22 MMT 21.09
Mangla 5 Dijabba Fault 8.78
Muzaffarabad 28 Riasi Thrust 1.78
Peshawar 20 MBT (west) 20.76
Talagang 30 Salt Range Thrust 23.56




23


Figure 3.1 Four steps of a deterministic seismic hazard analysis (Kramer 1996).


24


Figure 3.2 General tectonic setting of Pakistan with the velocities of plate movement from
Larson et al. (1999) and Bilham (2004). Major faults with two active belts are shown. The
dashed rectangle denotes the study area in Northwestern Pakistan.

~ 29 mm/year
Karachi
MBT : Main Boundary Thrust
MKT : Main Karakoram Thrust
MMT : Main Mantle Thrust
Baluchistan
Muzaffarabad
Islamabad
500 km
~ 44 mm/year
~ 42 mm/year
Pakistan
India
Afghanistan

25


Figure 3.3 Faults in NW Pakistan and earthquakes (M
w
4.0) from 1505 to 2006 used in this
study. Numbers beside faults for identification; see Table 3.4 for fault names and properties.
GPS measurements by Bendict et al. (2007) and inferred plate movements from Larson et al.
(1999) are shown as arrows. Zones 1, 2, and 3 were used to group faults to estimate individual
fault slip rates from available plate movement velocity data.


36 N
35 N
34 N
33 N
32 N
70 E 71 E 72 E 73 E 74 E 75 E
Astor
Kaghan
Malakand
Balakot
Muzaffarabad
Islamabad Kohat
Bannu
Talagang
Peshawar
GPS measurement by Bendick et al. (2007)
Inferred plate movement based on Larson et al. (1999)
Assumed plate movement
Earthquakes
M
w
4 5 6 7
Zone 1
Zone 2
Zone 3
Mangla
1914
1928
1974
1992
2002
2004
2005
2005
2005
1885

26


Figure 3.4 Definition of fault geometry parameters.

Dip
Depth to bottom
of rupture (D)
Depth to top
of rupture (Z
tor
)

27


Figure 3.5 PGA predictions for the five NGA GMPEs used in this study using M
w
= 7.6, Z
tor
= 0,
D = 15 km, and V
s,30
= 760 m/s for foot wall and hanging wall sides.
1 10 100
Distance from the fault (km)
0.01
0.1
1
P
e
a
k

G
r
o
u
n
d

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Foot wall side
AS (2008)
BA (2008)
CB (2008)
CY (2008)
I (2008)
Hanging wall side
AS (2008)
BA (2008)
CB (2008)
CY (2008)
I (2008)

28


Figure 3.6 Logic tree for faults in NW Pakistan (NWP) for use in the DSHA. The numbers in
parentheses represent the weights assigned to each branch. AS, BA, CB, I represent GMPEs
proposed by Abrahamson-Silva (2008), Boore-Atkinson (2008), Campbell-Bozorgnia (2008),
and Idriss (2008). The weights for the Z
tor
are shown in the table.





Depth to the
top of rupture
(Z
tor
)
0
2 km
4 km
10 (0.4)
15 (0.4)
20 (0.2)
0.8 (0.2)
0.9 (0.6)
1.0 (0.2)
NWP
Faults
Depth to the
bottom of
rupture (D)
Seismogenic
factor
(R-factor)
Attenuation
relationship
as above
as above
as above
AS (0.25)
BA (0.25)
CB (0.25)
I (0.25)
Magnitude
range
Z
tor
= 0 Z
tor
= 2 km Z
tor
= 4 km
6.5 s M s 6.75
6.75 < M s 7.0
7.0 < M
0.333 0.333 0.333
0.5 0.5
1.0

Weights for the depth to the top of rupture, Z


tor
Maximum
magnitude
Fault geometry

29


Figure 3.7 Sensitivity of PGA computed in DSHA to selected GMPEs.
0
0.2
0.4
0.6
0.8
1
1.2
1.4
P
G
A

(
g
)
Islamabad Muzaffarabad
Weighted
AS
BA
CB
I
Median
0
0.2
0.4
0.6
0.8
1
1.2
1.4
P
G
A

(
g
)
Islamabad Muzaffarabad
84
th
percentile

30


Figure 3.8 Contour maps of median PGA (left) and 84
th
percentile PGA (right) for the rupture
that occurred during the 2005 Kashmir earthquake using the DSHA procedure and V
s,30
= 760
m/s. A grid spacing of 0.1 0.1 was used for computations. For comparison to computed
PGAs, measurement stations are shown as black squares. The white line denotes the surface
rupture trace.


Kaghan
Balakot
Muzaffarabad
Islamabad
Kaghan
Balakot
Muzaffarabad
Islamabad
72.5E 73E 73.5E 74E
34.5N
35N
34N
33.5N
72.5E 73E 73.5E 74E
0.1 0.2 0.3 0.4 0.5 0.6 0.8 0.7 0.9 1.0 (g)
PGA

31


Figure 3.9 Comparison of PGA prediction using DSHA for (a) Site Class C (very dense soil and
soft rock, V
s,30
= 560 m/s) conditions and (b) Site Class D (stiff soil, V
s,30
= 270 m/s) conditions,
with strong ground motions (Durrani et al. 2005) measured during the 2005 Kashmir earthquake.
Two measurements for Abbottabad, Murree, and Nilore are two orthogonal recording directions.
Idriss (2008) GMPE does not apply for V
s,30
< 450 m/s. The measurement in Tarbela is suspected
to correspond to Site Class B.

1 10 100
Distance from the fault (km)
0.01
0.1
1
P
e
a
k

G
r
o
u
n
d

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Computed
Abrahamson-Silva (2008)
Boore-Atkinson (2008)
Campbell-Bozorgnia (2008)
Idriss (2008)
Mean of all median NGAs
Mean of all median NGAs 1 o
Measured
Abbottabad
Barotha
Murree
Nilore
Tarbela
1 10 100
Distance from the fault (km)
(a) V
s,30
= 560 m/s (b) V
s,30
= 270 m/s
200 200

32

CHAPTER 4. PROBABILISTIC SEISMIC HAZARD ANALYSIS
4.1 Introduction
The probabilistic seismic hazard analysis (PSHA) can incorporate uncertainties in
recurrence, M
max
, GMPE, and other parameters. The main benefit of PSHA is that it allows
computation of the mean annual rate of exceedance of a ground motion parameter at a particular
site based on the aggregate risk from potential earthquake sources.
Similar to DSHA, a typical PSHA consists of four main steps (Figure 4.1):
1) Identification and characterization of earthquake source zones. Uniform probability
distributions of potential rupture locations are assigned to each source.
2) Characterization of earthquake recurrence. A recurrence relationship, which specifies
the average rate at which an earthquake of some size will be exceeded, is developed.
3) Selection of attenuation relationships. The ground motion parameters at the site are
determined using the selected attenuation relationships.
4) Computation of ground motion parameters corresponding to a probability of
exceedance. The probabilities that the ground motion parameter will be exceeded
during a particular time period for all sources are combined.

For the PSHA, the same fault geometries and GMPEs used in the DSHA were
considered. However, PSHA also requires earthquake recurrence characteristics for each fault or
source. This chapter describes the methodology to characterize two common earthquake
recurrence: bounded exponential magnitude distribution and characteristic earthquake
distribution.
4.2 Bounded exponential distribution recurrence model
The exponential recurrence distribution was first introduced by Gutenberg and Richter
(1954), which can be expressed as:

bm a
m
= log Eq. (4-1)


33

where
m
is the complementary cumulative earthquake rate for magnitudes > m (i.e.,
m
is the
annual rate of exceedance of magnitude m), and a and b are constants. Eq. (4-1) can also be
expressed in exponential form as:

( ) m
m
| o = exp Eq. (4-2)

where = 2.303a and = 2.303b.
To limit the maximum earthquake magnitude, Cornell and Vanmarcke (1969) proposed
the bounded exponential recurrence, which can be expressed as:

( ) ( ) ( ) ( )
( ) ( )
0
0 0
exp 1
exp exp
m m
m m m m
u
u
m


=
|
| |
v for
u
m m s Eq. (4-3)

where ( )
0
exp m | o v = , m
0
is the threshold magnitude (m
0
= 4 in this study), and m
u
is the upper
bound magnitude. Thus m
0
, m
u
, , and are the important parameters that characterize the
recurrence relation for each fault.
4.2.1 Threshold magnitude, m
0
, and upper bound magnitude, m
u

The value of m
0
defines the starting point of the recurrence relationship. If m
0
were set
less than M
w
4.0, recurrence would be underestimated because the composite earthquake catalog
for NW Pakistan compiled for this study lacks data for M
w
< 4.0. Therefore, m
0
was set to be 4.0.
For the same reason, previous studies (NORSAR and PMD 2006; PMD and NORSAR 2006;
Monalisa et al. 2007) also used m
0
values of 4.0 to 4.5.
As noted previously, m
u
was set to be same as M
max
. The USGS (Petersen 2008) assigns
uncertainties in m
u
of 0.2 magnitude unit for epistemic uncertainty, and 0.24 magnitude unit for
aleatory uncertainty. Epistemic and aleatory uncertainties were combined in the analyses and a
total uncertainty of 0.5 magnitude units was applied to the mean upper bound magnitude.
4.2.2 Activity rate
Activity rate, in Eq. (4-3), is the complementary cumulative earthquake rate for m > m
0
.
The activity rate for each fault is calculated by dividing the cumulative number of earthquakes
occurring along each fault with M
w
> 4.0 (starting from large magnitude and working toward

34

smaller magnitude) by the corresponding time interval (T
s
). Earthquakes with large focal depth
were excluded from the analysis unless the magnitude was sufficiently large to influence the
seismic hazard. Specifically, earthquakes with focal depths > 35 km and M
w
< 5, as well as those
with focal depth > 45 km and M
w
< 6, were excluded. Two estimates of activity rate (as
explained below) were computed for each fault as shown in Table 3.4.
Figure 4.2 presents the instrumented earthquake distribution in NW Pakistan from 1960
to 2006. As discussed earlier, it is clear that the catalog is incomplete, at least for times prior to
about 1972. Therefore, using the sample from entire time interval severely underestimates the
mean rate of occurrence. On the other hand, a reduced time interval may not be suitable to define
return periods of large earthquakes. Therefore, it is important to determine the time interval
where the mean complementary cumulative earthquake rate () remains unchanged over time.
The completeness analysis proposed by Stepp (1972) was used for three magnitude groups
(4M
w
<5, 5M
w
<6, and M
w
6). Stepp (1972) suggested that if the slope of the standard
deviation of (
s
T o

= ) is parallel to the slope of


s
T 1 , can be considered as stable for
that T
s
. Figure 4.3 shows the result of completeness analysis for time intervals that increase in 5-
year steps, measured from 2006 (i.e., 2002-2006, 1997-2006, 1992-2006, etc.). The events for 4
M
w
< 6 are complete during the 5- to 30-year interval. The departure of

from
s
T 1 after
30-year interval can be explained by incomplete reporting of earthquakes. The events for M
w
6
are complete during the 55- to 100-year interval. The large earthquakes (M
w
6) are much fewer
than small and intermediate earthquakes (4M
w
<6) and do not significantly influence the
estimate of activity rate. In addition, the

for these large earthquakes is slightly higher than, and


parallel to the line of
s
T 1 during the 25- to 30-year interval. Therefore, Activity rate 1 was
computed for each fault using a 25-year time interval in the earthquake catalog from 1981 to
2006 (Period 1) where for 4M
w
<6 seems reasonably stable.
Figure 4.2 clearly shows a spike in the instrumented catalog in 2005, associated with the
large number of foreshocks and aftershocks (M
w
4) accompanying the 2005 Kashmir
earthquake. Therefore, it is suspected that the foreshocks and aftershocks associated with this
large event could bias the hazard calculation. Thus, Activity rate 2 was computed for each fault
using a 30-year period in the catalog from 1975 to 2004 (Period 2), excluding earthquakes
associated with the 2005 Kashmir event. Figure 4.4 compares activity rates 1 and 2 for the faults.

35

Then, the earthquakes from 1981 to 2006 were declustered using the method by Gardner and
Knopoff (1974). The total activity rate using these declustered earthquakes was estimated to be
approximately 15.2 per year which is between two values of activity rate using Period 1 (30.7 per
year) and Period 2 (9.4 per year). Therefore, it can be concluded that the approach using two
activity rates is reasonable.
4.2.3 b-value
The b-value establishes the slope of the exponential recurrence model. Figure 4.5 shows
the regression lines using Least Square (LS) method and Maximum Likelihood (ML) method
(Weichert 1980) for the complementary cumulative rate of observed earthquakes from Period 1
(1981-2006) and Period 2 (1975-2004) for 4M
w
<6. Figure 4.5 also shows b = 0.8 lines, which is
the b-value used by most faults in the USGS seismic hazard mapping project (Petersen 2008).
These comparisons illustrate that the regression lines using LS and ML methods poorly fit the
data through all magnitude ranges. For earthquakes occurring during Period 1 (1981-2006) in
Figure 4.5 (a), the LS line captures the overall trend of earthquake rate, but it significantly
underestimates the complementary cumulative earthquake rate for the largest magnitude. On the
other hand, the ML line captures the data at small magnitude range and the largest magnitude.
The b = 0.8 line is similar to the ML line.
For earthquakes for Period 2 (1975-2004) in Figure 4.5 (b), the ML line fails to capture
the data above M
w
= 4.5, while the LS line fits most of data. The b = 0.8 line is similar to the LS
trend line. For declustered earthquakes (1981-2006) in Figure 4.5 (c), the ML and LS lines are
similar and fit most of data except for the largest magnitude. The b = 0.8 line is similar to ML
and LS lines, but closer to the point for the largest magnitude, maintaining the best fit for small
magnitude range. Considering these comparisons, it was concluded that the b = 0.8 line provides
a reasonable fit to the earthquake data (especially for small and large magnitudes), and this b-
value was used for all faults in NW Pakistan. Varying the b-value from 0.80 to 0.85 does not
significantly affect the PGA. For example, the PGAs with 475-year return period for
Muazaffarabad using b = 0.80 and 0.85 were estimated to be 0.64g and 0.62g, respectively.


36

4.3 Characteristic distribution recurrence model
Youngs and Coppersmith (1985) observed that large magnitude earthquakes commonly
occur at a higher rate than predicted by the exponential distribution model (Figure 4.6). These
large earthquake are called the characteristic earthquake. They proposed an alternate recurrence
model, termed the characteristic distribution model, that combines an exponential magnitude
distribution up to magnitude m' with a uniform distribution in the magnitude range of m
u
- m
c
to
m
u
at a rate density ( )
c
m n . The present study used the characteristic model with characteristic
magnitude range, m
c
, of 0.5 as suggested by Youngs and Coppersmith (1985). The
characteristic earthquake, ( )
c
m N

, was estimated by various methods as explained subsequently.


The upper bound magnitude was used as the characteristic magnitude for each fault. A total
uncertainty of 0.5 magnitude units was applied to the mean characteristic magnitude.
Characteristic rate is the rate at which the characteristic earthquake occurs and is usually
determined using the slip rate estimated by geodetic and geologic evidence, cosmogenic dating
methods using
14
C,
3
He, or
10
Be, and paleoseismic data for the region. However, as discussed
above, faults in NW Pakistan are not thoroughly investigated and the slip rate of each fault is not
accurately known. In addition, the historical seismicity record is not long enough to include the
characteristic events. Therefore, it is difficult to estimate the characteristic rate for faults in NW
Pakistan. As a result, Kim et al. (2010) only employed exponential recurrence models to evaluate
seismic hazard in Islamabad and Muzaffarabad. This study proposes three alternative methods to
estimate characteristic rates. These three alternative characteristic rates for mean characteristic
magnitude are listed for each fault in Table 3.4.
4.3.1 Characteristic rate A (relation between slip rate and characteristic rate)
The slip rates and characteristic rates for 250 faults in the U.S. Intermountain West, 60
faults in the U.S. Pacific Northwest, and 151 A-type and B-type faults in California are compared
in Figure 4.7. The two regression fits were tested as shown in Figure 4.7. Although Fit 2 is
simpler than Fit 1, the coefficient of determination (R
2
) for Fit 2 is smaller than that for Fit 1,
suggesting that Fit 1 fits data better than Fit 2. Therefore, Fit 1 was selected in this study, which
is expressed as:

( ) ( )
8988 . 0
0008016 . 0 rate Slip rate stic Characteri = Eq. (4-4)

37

Bendick et al. (2007) reported that velocities inferred from GPS measurements from 2001
to 2003 for observation sites near Balakot and Islamabad are approximately 52 mm/yr and 38
mm/yr, respectively. It was assumed that the velocity of the Indian Plate in the southern part of
the study area and that of the Eurasian Plate in the northern part of the study area are the same as
the values observed by Larson et al. (1999) (44 mm/year and 29 mm/year, respectively). The
difference in velocities (52 mm/year 29 mm/year) yields a total compression rate for Zone 1 of
approximately 23 mm/year. Similarly, a total extension rate for Zone 2 of about 14 mm/yr can be
computed from the difference in velocities of 52 mm/year and 38 mm/year. For Zone 3, the
shape of the Khair-I-Murat fault suggested that it is a normal fault formed by differential slip on
its northern and southern segments. The MBT west segment is a thrust fault where considerable
amount of compression stress occurs. The Khairabad fault, next to the MBT west segment, is
also considered as reverse fault. Based on these considerations, the velocity of 44 mm/year was
assumed on the southern part of MBT, which results in compressional slip rates of 4 mm/year for
the MBT west segment and 2 mm/year to Khairabad fault. Another velocity of 40 mm/year was
assumed on the southern parts of the Khair-I-Murat fault. This assumed velocity results in an
extentional slip rate of 4 mm/year for the Khair-I-Murat fault . The total slip rate for the
remaining faults in Zone 3 is then 4 mm/year based on the difference in the measured velocity
(44 mm/year) and the assumed velocity (40 mm/year). The total slip rate for each zone was
distributed to individual faults by assuming that slip rate is proportion to fault length.
The assigned slip rate was used with Eq. (4-4) to compute the characteristic rate for the
mean characteristic magnitude for each fault. The characteristic rates for m
c
0.5 magnitude
units was computed using the relationship between return period and magnitude proposed by
Slemmons (1982) as shown below.

( )
47 . 0
5 . 0
=
+
c
c
m for rate stic Characteri
m for rate stic Characteri

Eq. (4-5 a)


( )
12 . 2
5 . 0
=

c
c
m for rate stic Characteri
m for rate stic Characteri

Eq. (4-5 b)

4.3.2 Characteristic rate B (relation between activity rate and characteristic rate)

38

The second method is to assume that the characteristic rate is proportional to the activity
rate of each fault. To estimate the overall characteristic rate, earthquakes with M
w
> 6.5 over a
100-year period (1907-2006) for NW Pakistan were considered. The completeness analysis
(Figure 4.3) indicated that the earthquake catalog compiled in this study provides a reasonably
complete representation of moderate to large earthquakes (M
w
6.0) over this 100-year time
frame. The measured earthquake rate for M
w
> 6.5 was estimate to be approximately 0.03 (per
year) as shown in Figure 4.8. This earthquake rate then was distributed to each fault proportional
to its activity rate. Because two different activity rate sets (Activity rate 1 and Activity rate 2)
were used, there are also two characteristic rate B sets. The average of the two sets was used in
the analysis.
4.3.3 Characteristic rate C (seismic moment balance method)
Many researchers (WGCEP 2003; Petersen 2008; WGCEP 2008) use the seismic
moment balance method to estimate the characteristic rate of faults. Aki (1966) proposed a
relationship between seismic moment, M
0
, and average fault rupture displacement per event
over the slip surface, d.

( )d RA M =
0
Eq. (4-6)

where is fault rigidity (usually taken to be 310
11
dyne/cm
2
).
The average displacement per event can be also expressed as S T
s
where S is the
average slip rate and T
s
is the recurrence interval (Wallace 1970). Seismic moment can be
estimated as a function of earthquake magnitude using the equation below proposed by Hanks
and Kanamori (1979).

1 . 16 5 . 1 log
0
+ =
w
M M Eq. (4-7)

Using the characteristic magnitude (defined earlier) to compute M
0
, the characteristic rate
can be calculated as:
( )
( )
0
M
S RA
rate istic Characeter

=
Eq. (4-8)

39


4.4 Total recurrence
Using all the parameters mentioned above, recurrence models for each fault were
established. Figure 4.8 presents the total recurrence models (exponential and characteristic
models), i.e., the sum of all recurrence models for individual faults in NW Pakistan. The
weighted average of activity rates, characteristic rates, and R-factors were used to construct this
total recurrence. The complementary cumulative rate of earthquake events for various time
intervals and regions were compared with the total recurrence as shown in Figure 4.8. In the
small magnitude range, the total complementary cumulative rates of the exponential models are
comparable to the observed earthquake rates. They slightly overestimate the middle magnitude
range, but fit well the data for magnitude around 6.5 M
w
. This phenomenon can be also observed
in the fault recurrence model for Northern California (WGCEP 2008). The total complementary
cumulative rates for large magnitude range (M
w
> 6.5) computed based on 25- to 30-year interval
are also consistent with observed earthquake rate using 100-year interval. Therefore, exponential
rates appear to be reasonable averages between data in the short interval and data in the long
interval.
The measured earthquake rates for large earthquakes for all of Pakistan and for only NW
Pakistan with two different time intervals (1907-2006 and 1505-2006) are included in Figure 4.8.
The maximum earthquake magnitudes in NW Pakistan in both time periods of 1907-2006 and
1505-2006 match with the mean upper bound magnitude 0.5 magnitude unit. The maximum
earthquake magnitude throughout all of Pakistan matches with the mean upper bound magnitude
+ 0.5 magnitude unit. The predicted earthquake rates from both exponential and characteristic
models are comparable to the measured earthquake rates in NW Pakistan in time period of 1907-
2006 and throughout all of Pakistan in time period of 1505-2006. Overall, the total recurrence
model is in good agreement with the seismicity data in Pakistan.
4.5 Logic trees for PSHA
Figure 4.9 shows the logic tree used for the PSHA. The PSHA logic tree contains the
same branches for fault geometry, R-factor, and attenuation relationships as that used in the
DSHA. A weight of 0.6 was assigned to the mean upper bound magnitude and 0.2 to the mean
upper bound magnitude 0.5 units. For the recurrence models, the characteristic model was

40

weighted at 2/3 and bounded exponential model was weighted at 1/3. These weights were
selected because there is evidence that the characteristic recurrence relationship captures the
earthquake rate of individual faults better than the bounded exponential recurrence relationship
(Youngs and Coppersmith 1985). The two sets of activity rates for the bounded exponential
recurrence model were equally weighted. Weights of 0.25 were assigned to the proposed
Characteristic rates A and B, and a weight of 0.5 was assigned to Characteristic rate C, the
commonly-used moment balance method.
4.6 PSHA sensitivity analysis
Sensitivity analyses for two parameters in the PSHA logic tree, earthquake rate and upper
bound magnitude, were performed to evaluate the influence of these parameters on the computed
seismic hazard. The sensitivity analysis for earthquake rate is shown Figure 4.10. The results are
shown for Islamabad and Muzaffarabad for a 2475-, 975-, and 475-year return periods. For
Islamabad, the PSHA yielded the highest PGA using characteristic rate A. The PGA estimated
using activity rate 1 is significantly high for Muzaffarabad, while the PGA estimated using
activity rate 1 is almost the same as that estimated using activity rate 2 for Islamabad. Activity
rate 1 includes the numerous aftershocks of 2005 Kashmir earthquake. Therefore, activity rate 1
for the faults near Muzaffarabad (located near the 2005 fault rupture) is greater than activity rate
2. In contrast, there is no significant difference between activity rate 1 and activity rate 2 for
faults near Islamabad.
The sensitivity of PGA to the mean upper bound magnitude is shown in Figure 4.11. It is
notable that the greater magnitudes yield lower PGAs because the characteristic rates were
adjusted according to the magnitude. That is, smaller characteristic rates were assigned to larger
magnitudes because larger magnitude earthquakes tend to occur less frequently. The weighted
average PGA is close to that computed using the mean magnitude value.
4.7 Evaluation of the proposed PSHA procedure
Figure 4.12 shows contour maps of PGA for 475- and 2475-year return periods over the
same area evaluated using DSHA. The proposed PSHA procedure was used to generate this
contour map, and PGAs are reported for V
s,30
= 760 m/s. For Muzaffarabad, the PGAs
corresponding to the 475- and 2475-year return periods are approximately 0.6g and 1.0g,
respectively, which are consistent with median PGA and 84
th
percentile PGA, respectively,

41

obtained from the DSHA of 2005 Kashmir earthquake (Section 3.9). The PSHA and DSHA
results compare similarly for Balakot.
4.8 Summary and discussion
This chapter describes the methodology to characterize the earthquake recurrence models
for faults in NW Pakistan. The important parameters that characterize the bounded exponential
recurrence model are m
0
, m
u
, and b-value. If m
0
were set less than M
w
4.0, recurrence would be
underestimated because the composite earthquake catalog for NW Pakistan compiled for this
study lacks data for M
w
< 4.0. Therefore, m
0
= 4.0 was used. The M
max
computed based on the
rupture area were used as m
u
. The completeness analysis revealed that earthquake catalog for 25-
to 30-year time interval from 2006 is reasonably complete. Therefore, the Activity rate 1 for
each fault was computed using a 25-year time interval in the earthquake catalog from 1981 to
2006 (Period 1). However, large number of foreshocks and aftershocks accompanying the 2005
Kashmir earthquake were suspected to bias the seismic hazard calculation. Thus, Activity rate
2 for each fault was computed using a 30-year period in the catalog from 1975 to 2004 (Period
2), excluding earthquakes associated with the 2005 Kashmir event. These activity rates were
compared well with the one computed by using declustered earthquakes. It was found that the
regression lines using least square and maximum likelihood methods could not fit the overall
trend of complementary cumulative earthquake rate. Therefore, b = 0.8, used or most faults in
the USGS seismic hazard mapping project (Petersen 2008), was chosen because it provide more
reasonable fit to the earthquake data.
For characteristic recurrence model, the upper bound magnitude was used as the
characteristic magnitude for each fault. Three methods to estimate characteristic rates are
described in this chapter. The first method assumes that the characteristic rate is related with the
slip rate of faults. The second method employs the assumption that the characteristic rate is
proportional to the activity rate. The third method is using the seismic moment balance method
proposed by Aki (1966). All these parameters and models were implemented in the logic tree and
reasonable weight for each branch of the logic tree was assigned. Sensitivity analyses for
earthquake rate and upper bound magnitude were performed to evaluate the influence of these
parameters on the computed seismic hazard. The PGA estimated using Activity rate 1 is
significantly high for Muzaffarabad due to inclusion events from the 2005 Kashmir earthquake,
while the PGA estimated using Activity rate 1 is almost the same as that estimated using Activity

42

rate 2 for Islamabad. Using the methodology developed in this chapter, the PGAs corresponding
to the 475- and 2475-year return periods are approximately 0.6g and 1.0g, respectively, which
are consistent with median PGA and 84
th
percentile PGA, respectively, obtained from the DSHA
of 2005 Kashmir earthquake.


43


Figure 4.1 Four steps of a probabilistic seismic hazard analysis (Kramer 1996).


44


Figure 4.2 Earthquake distribution (31.5N-36N, 69.5E-75.5E) as a function of time. Two
different periods were considered to estimate earthquake recurrence for the faults considered in
this study. Period 1 is from 1981 to 2006 and the Period 2 is from 1975 to 2004.


Figure 4.3 Completeness analysis for NW Pakistan for different magnitudes using Stepp (1972)
method.

= standard deviation, = rate of earthquake occurrence, and T


s
= time interval. The
solid lines are parallel to
s
T 1 .
1960 1965 1970 1975 1980 1985 1990 1995 2000 2005
Time
(year)
0
100
200
300
400
500
600
700
N
u
m
b
e
r

o
f

e
v
e
n
t
Period 1
Period 2
1 10 100
Time interval, T
s
(years)
0.01
0.1
1
10
o


=

\

T
s
4s M
w
<5
5s M
w
<6
M
w
>6

45


Figure 4.4 Comparison of fault activity rates (Activity rate 1:1982 to 2006; and Activity rate 2:
1975 to 2004). Numbers represent the fault number used in this study (see Table 3.4).

0.01 0.1 1 10
Activity rate 1 (year
-1
)
0.01
0.1
1
10
A
c
t
i
v
i
t
y

r
a
t
e

2

(
y
e
a
r
-
1
)
2
3
5
6
7
8
9
10
11
12
13
14 15
16
17
18
20
21
22
23
24
25
26
28
29
30
31
32
1
1

46


Figure 4.5 Regression lines using Least Square (LS) method and Maximum Likelihood method
(Weichert 1980) for the complementary cumulative rate of observed earthquakes for (a) Period 1
(1981-2006), (b) Period 2 (1975-2004), and (c) declustered catalog (1981-2006). The b = 0.8 line
starting from an annual earthquake rate at M
w
=4 for each case is also shown.




4 4.5 5 5.5 6 6.5 7 7.5 8
Magnitude, M
w
0.01
0.1
1
10
100
A
n
n
u
a
l

r
a
t
e

o
f

e
v
e
n
t
s

>

M
w

(
y
e
a
r
-
1
)
Least Sqare (LS) method
Maximum Likelihood (ML) method
Line with b = 0.8
4 4.5 5 5.5 6 6.5 7 7.5 8
Magnitude, M
w
4 4.5 5 5.5 6 6.5 7 7.5 8
Magnitude, M
w
0.01
0.1
1
10
100
A
n
n
u
a
l

r
a
t
e

o
f

e
v
e
n
t
s

>

M
w

(
y
e
a
r
-
1
)
(a)
(b)
(c)
b
ML
= 0.82
b
LS
= 0.96
b
ML
= 1.17
b
LS
= 0.74
b
ML
= 0.92
b
LS
= 0.85

47


Figure 4.6 Hypothetical recurrence relationship (Youngs and Coppersmith 1985).

48


Figure 4.7 Relationship between slip rates and characteristic rates of western U.S faults. A-type
faults in California are divided into several segments, each with a unique slip rate. The
characteristic rates for these faults were calculated using three methods [a-priori, Ellsworth of
WGCEP (2003), and Hanks and Bakun (2008)] used by WGCEP (2008). Error bars are used to
illustrate uncertainties in slip rates and characteristic rates for these faults.

0.001 0.01 0.1 1 10 100
Slip rate (mm/yr)
1E-006
1E-005
0.0001
0.001
0.01
0.1
C
h
a
r
a
c
t
e
r
i
s
t
i
c

r
a
t
e

(
y
e
a
r
-
1
)
California faults (A-Type)
California faults (B-Type)
Intermountain West faults
Pacific Northwest faults
Fit 1
Fit 2
: Characteristic rate = 0.0008016 Slip rate
0.8988
(R
2
= 0.9038)

: Characteristic rate = 0.001 Slip rate (R
2
= 0.6427)


49


Figure 4.8 Complementary cumulative rate of observed earthquakes and recurrence models from
the composite catalog used in this paper. The total recurrence models represent the sum of the
recurrence models of all of the individual faults. The recurrence models for the east and west
MBT segments are shown as examples.


4 4.5 5 5.5 6 6.5 7 7.5 8 8.5 9
Magnitude, M
w
1E-005
0.0001
0.001
0.01
0.1
1
10
100
A
n
n
u
a
l

r
a
t
e

o
f

e
v
e
n
t
s

>

M
w

(
y
e
a
r
-
1
)
Earthquakes in NW Pakistan during Period 1 (1981-2006)
Earthquakes in NW Pakistan during Period 2 (1975-2004)
Declustered earthquakes in NW Pakistan (1981-2006)
Large earthquakes (M
w
>6.5) in all of Pakistan (1907-2006)
Large earthquakes (M
w
>6.5) in all of Pakistan (1505-2006)
Large earthquakes (M
w
>6.5) in NW Pakistan (1907-2006)
Large earthquakes (M
w
>6.5) in NW Pakistan (1505-2006)
Total exponential model with mean M
max
and mean M
max
0.5
Total characteristic model with mean M
max
and mean M
max
0.5
MBT exponential model with mean M
max
MBT characteristic model with mean M
max
M
B
T
(
w
e
s
t
)
M
B
T

(
e
a
s
t
)

50


Figure 4.9 Logic tree for faults in NW Pakistan used in the PSHA. Numbers in parentheses
represent weights assigned to each branch. The weights for Z
tor
are the same as used for the
DSHA (Figure 3.6).
Depth to the
top of rupture
(Z
tor
)
0
2 km
4 km
10
(0.4)
15
(0.4)
20
(0.2)
0.8
0.9
(0.2)
(0.6)
1.0
(0.2)
Mean + 0.5
(0.2)
Mean
(0.6)
Mean - 0.5
(0.2)
Exponential
(0.33)
Charact-
eristic
(0.67)
Activity rate 1
(0.5)
Activity rate 2
(0.5)
Charcteristic
rate A
(0.25)
Characteristic
rate B
(0.25)
Characteristic
rate C
(0.5)
AS
(0.25)
BA
(0.25)
CB
(0.25)
I
(0.25)
NWP
Faults
Depth to the
bottom of
rupture (D)
Seismogenic
factor
(R-factor)
Magnitude
uncertainty
Recurrence model
Earthquake
rate
Attenuation
relationship
Maximum magnitude Fault geometry
as above
as above
as above
as below
as below

51


Figure 4.10 Sensitivity of PGA computed at Islamabad and Muzaffarabad in the PSHA to the
characteristic rates and activity rates for return periods of (a) 2475 years (b) 975 years, and (c)
475 years. All other parameters are used with weights shown in the logic tree (Figure 4.9).

0
0.2
0.4
0.6
0.8
1
1.2
1.4
P
G
A

(
g
)
0
0.2
0.4
0.6
0.8
1
1.2
1.4
P
G
A

(
g
)
0
0.2
0.4
0.6
0.8
1
1.2
1.4
P
G
A

(
g
)
Islamabad Muzaffarabad
Weighted average
Char.A
Char.B
Char.C
Act.1
Act.2
(a)
(b)
(c)

52


Figure 4.11 Sensitivity of PGA computed at Islamabad and Muzaffarabad in the PSHA to the
upper bound magnitudes for return periods of (a) 2475 years (b) 975 years, and (c) 475 years.
All other parameters are used with weights shown in the logic tree (Figure 4.9).

0
0.2
0.4
0.6
0.8
1
1.2
1.4
P
G
A

(
g
)
0
0.2
0.4
0.6
0.8
1
1.2
1.4
P
G
A

(
g
)
0
0.2
0.4
0.6
0.8
1
1.2
1.4
P
G
A

(
g
)
Islamabad Muzaffarabad
Weighted average
Mean + 0.5
Mean
Mean - 0.5
(a)
(b)
(c)

53


Figure 4.12 Contour maps of PGA for 475- (left) and 2475-year (right) return periods using the
PSHA procedure and V
s,30
= 760 m/s. A grid spacing of 0.1 0.1 was used for computations.


Kaghan
Balakot
Muzaffarabad
Islamabad
Kaghan
Balakot
Muzaffarabad
Islamabad
72.5E 73E 73.5E 74E
34.5N
35N
34N
33.5N
72.5E 73E 73.5E 74E
0.8
0.9
1
.
0
0.7
0.7
0
.
6
0.1 0.2 0.3 0.4 0.5 0.6 0.8 0.7 0.9 1.0 (g)
PGA

54

CHAPTER 5. SEISMIC HAZARD RESULTS
This chapter summarizes the results of both probabilistic and deterministic seismic
hazard analyses performed for the 11 cities for an assumed bedrock with V
s,30
of 760 m/s. (Site
effects can be considered separately by performing site response analysis or by applying site
coefficients.), using the procedures described in Chapter 3 and Chapter 4. This chapter presents
PSHA hazard curves for PGA, PSHA uniform hazard spectra (UHS), and DSHA hazard spectra
for the 11 cities and seismic hazard maps of PGA and SAs (T=0.2 and 1.0 sec) for 475-, 975-,
and 2475-year return periods. Based on deaggregation results, a discussion of the conditional
mean spectra for engineering applications is also presented.
5.1 Islamabad
Figure 5.1 presents the seismic hazard curves for PGA at 5% damping and select uniform
hazard response spectra for the city of Islamabad. Hazard curves from the top five contributing
faults are included in the figure. The PGAs corresponding to 475-, 975-, and 2475-year return
periods are approximately 0.35g, 0.48g, and 0.69g, respectively. The individual fault hazard
curves illustrate that the MBT west segment contributes the most to the hazard at most
commonly-employed return periods. This fault is located closer to the city than any other fault,
exhibits a high activity rate, and has a large maximum moment magnitude, M
max
. Response
spectra (Figure 5.1b) can be used to evaluate structural response at different natural periods. The
SA is highest at T = 0.2 sec (1.73g for a 2475-year return period), while at T = 1.0 sec, SA is
estimated to be 0.67g for 2475-year return period, which is similar to the PGA value.
Deaggregation of seismic hazard for Islamabad was conducted for three SAs (at T = 0.05
sec, 0.2 sec, and 1.0 sec) for a 475-year return period as shown in Figure 5.2. Deaggregation
provides the relative contribution to seismic hazard from each fault (i.e., seismic source) in terms
of M
w
, source-to-site distance, R, and ground motion uncertainty, . The magnitude bin width is
0.2 M
w
units, and the distance bin width is 5 km units. The mean and modal values of M
w
, R, and
that correspond to a given maximum amplitude can be calculated from the deaggregation
analysis. Mean values of M
w
and R are commonly used to represent the controlling earthquake
size and location for developing site-specific time histories (Bernreuter 1992; U.S. Department
of Energy (U.S. DOE) 1996; U.S. Nuclear Regulatory Commission (U.S. NRC) 1997).
The mean values of M
w
, R, and for Islamabad are shown in Table 5.1. These mean
values of M
w
and R can be used to develop conditional mean spectra (discussed in a later

55

section), perform liquefaction analyses, or analyze seismic slope stability, etc. Modal values are
used to select the controlling earthquake when a site is proximate to two equally hazardous faults
(Bazzurro and Cornell 1999). For Islamabad, large earthquakes on the nearby MBT west
segment dominate the hazard. There is a small contribution from more distant faults for T = 1.0
sec. Deaggregation analyses for 975-and 2475-year return periods show similar trends.
A DSHA spectrum (Figure 5.1b) was constructed for the controlling fault (MBT west
segment) determined by deaggregation. The median DSHA spectrum is comparable to the UHS
for a 975-year return period, while the 84
th
percentile DSHA spectrum is higher than the UHS for
a 2475-year return period.
5.2 Muzaffarabad
The seismic hazard results for the city of Muzaffarabad are shown in Figure 5.3. As
shown in the PSHA hazard curve, PGAs corresponding to 475-, 975-, and 2475-year return
periods are 0.64g, 0.80g, and 1.02g, respectively. These PGAs are greater than those for
Islamabad because Muzaffarabad is proximate to more faults with higher activity rates and larger
M
max
. It can be noted that the Jhelum fault contributes most to the hazard. The highest spectral
acceleration is 2.70g at a period of 0.2 sec for a 2475-year return period.
The deaggregation results for the three SAs (at T = 0.05 sec, 0.2 sec, and 1.0 sec) for a
475-year return period for Muzaffarabad are shown in Figure 5.4. Similar to Islamabad, large
earthquakes on the nearby faults dominate the hazard at all return periods and spectral periods.
There is a small contribution from more distant faults for T = 1.0 sec. Mean values of M
w
, R, and
for Muzaffarabad are shown in Table 5.1.
DSHA spectra were also constructed for Muzaffarabad using the controlling Riasi fault.
Overall, the median DSHA spectrum is similar to the 475-year return period UHS for T > 0.5 sec,
but lower than the 475-year return period UHS for T < 0.5 sec, while the 84
th
percentile DSHA
spectrum ranges from the 975-year return period at short periods to the 2475-year return period
UHS at long periods.
5.3 Kaghan
The seismic hazard results for the city of Kaghan are shown in Figure 5.5. The PGAs for
475-, 975-, and 2475-year return periods are estimated to be 0.58g, 0.75g, and 1.00g,
respectively. These values are comparable with those for Muzaffarabad. The MCT fault

56

contributes significantly to the hazard at PGAs greater than 0.4g. For smaller PGAs, the MMT
and the Balakot Shear Zone contribute most to the hazard. The UHS for the 2475-year return
period yields SAs of 2.67g and 1.06g at T = 0.2 sec and 1.0 sec, respectively. The 84
th
percentile
DSHA spectrum is in good agreement with UHS for a 2475-year return period, while the median
DSHA spectrum falls between the UHS for 475- and 975-year return periods.
5.4 Peshawar
The PGAs for 475-, 975-, and 2475-year return periods for the city of Peshawar were
estimated to be 0.21g, 0.28g, and 0.38g, respectively, from the hazard curve in Figure 5.6. These
PGAs are considerably smaller than those for Islamabad, Muzaffarabad, and Kaghan mainly
because of the larger source-to-site distances for Peshawar. The MBT west segment and
Hissartang fault are the two primary contributing faults (Figure 5.6a). The MBT west segment
contributes slightly more to hazard despite a larger source-to-site distance because it exhibits a
larger M
max
and higher activity rate than the Hissartan fault. The SAs at T = 0.2 sec and 1.0 sec
for a 2475-year return period are 1.01g and 0.42g, respectively. The median and 84
th
percentile
DSHA spectra are comparable to the UHS for 475- and 2475-year return periods, respectively.
5.5 Summary of seismic hazard analyses for eleven NW Pakistan cities
Figure 5.7 compares SAs at T = 0.01 sec (PGA), 0.2 sec, and 1.0 sec computed using
both DSHA and PSHA for the 11 cities in NW Pakistan shown in Figure 1 for an assumed
bedrock outcrop with V
s,30
= 760 m/s. The highest PGAs are estimated to occur in Kaghan and
Muzaffarabad, with PGAs for a 2475-year return period exceeding 1.0g as a result of their
proximity to major active faults. Although Islamabad and Kohat are both approximately 5 km
away from the MBT west segment, the PGAs for these cities are not as high as Kaghan and
Muzaffarabad because they are located on the foot-wall side of the fault. The cities of Astor,
Bannu, Malakand, Mangla, Peshawar, and Talagang are located much farther from significant
faults; therefore, PGAs for these cities are considerably lower. Accelerations estimated by DSHA
are comparable to those from PSHA. In general, median accelerations agree well with those
corresponding to 475- or 975-year return periods, while the 84
th
percentile accelerations are
comparable to those corresponding to the 2475-year return period. The PSHA hazard curves,
PSHA uniform hazard spectra, and DSHA hazard spectra for the rest of cities are shown in
Figure 5.8 through Figure 5.14.

57

5.6 Comparison with previous studies
Figure 5.15 compares PGAs corresponding to a 475-year return period from PSHA and
median PGAs from DSHA with those computed in previous studies (PMD and NORSAR 2006;
Monalisa et al. 2007). This comparison clearly illustrates the differences between computed
seismic hazard using individual faults and that using area sources or diffuse seismicity for cities
close to active faults (e.g., Islamabad, Kaghan, Kohat, and Muzaffarabad). However, the
differences are smaller for cities located far from active faults. For example at Astor, Monalisa et
al. (2007) and PMD and NORSAR (2006) computed PGAs of 0.08g for a 475-year return period
and 0.28g for a 500-year return period, respectively. This study yielded similar PGAs: 0.21g for a
475-year return period from PSHA, and median PGA of 0.21g from DSHA because only a few
faults exist near this city and the distance between controlling fault (the MMT) and the city is
approximately 24 km.
Monalisa et al. (2007) and PMD and NORSAR (2006) both predicted the smallest PGA
for Bannu (0.07g and 0.08g, respectively). Although the faults near this city have lower activity
rates compared to other faults in NW Pakistan, these PGA values appear too small because the
city is located less than 10 km from an active fault. In contrast, a PGA of 0.16g was computed
for a 475-year return period from the PSHA and a median PGA of 0.30g from the DSHA.
Similarly, Islamabad is located less than 5 km from the MBT west segment, one of the most
hazardous faults in NW Pakistan. Previous studies predicted very small PGAs for this city: 0.13g
by Monalisa et al. (2007) and 0.20g by PMD and NORSAR (2006). In contrast, for Islamabad a
PGA = 0.35g was computed for a 475-year return period using PSHA and a median PGA = 0.51g
using DSHA.
As additional examples, Monalisa et al. (2007) and PMD and NORSAR (2006) computed
PGAs of 0.11g and 0.20g for Kaghan, respectively, despite the city being surrounded by several
active faults with high activity rates. In contrast, the highest PGAs were computed for this city,
with a PGA = 0.58g for 475-year return period (PSHA) and a median PGA = 0.72g (DSHA).
Similarly, Monalisa et al. (2007) and PMD and NORSAR (2006) computed PGAs of 0.21g and
0.13g, respectively, for Kohat. However, this city, like Islamabad, is located adjacent to the MBT
west segment. As a result, this study computed much larger PGAs, similar to those for Islamabad.
Lastly, Muzaffarabad is located close to the epicenter of the 2005 Kashmir earthquake,
and experienced severe damage during that earthquake. Moreover, this city is proximate to

58

several active faults including the MBT(east), MCT, and Jhelum fault. Despite this, Monalisa et
al. (2007) and PMD and NORSAR (2006) computed PGAs of 0.15g and 0.20g, respectively, for
Muzaffarabad. In contrast, this study computed much larger PGAs: 0.64g with 475-year return
period by PSHA and median PGA of 0.54g by DSHA that compare reasonably with those
measured in 2005. These values are also comparable with the deterministic prediction (PGA =
0.66g) by Durrani et al. (2005), using the GMPE by Ambraseys et al. (2005a). Only in cities
located farther from mapped faults (e.g., Malakand, Mangla, Peshawar, and Talagang) do the
PGAs computed in this study compare reasonably with those computed by Monalisa et al. (2007)
and PMD and NORSAR (2006).
In summary, PGAs estimated in this study using individual faults show distinct
differences at cities located at variable source-to-site distances, in contrast to previous studies
that used areal source zones, and computed PGAs that rarely exceeded 0.2g and showed little
variation among cities. As anticipated, PGAs computed for cities located far from faults (e.g.,
Astor, Malakand, Mangla, Peshawar, and Talagang) are similar to or slightly higher than those
computed using areal source zones. In contrast, PGAs computed for cities located close to faults
(e.g., Islamabad, Kaghan, Kohat, and Muzaffarabad) are 2 to 4 times greater than those predicted
by previous studies using areal source zones.
5.7 Hazard maps
Seismic hazard maps for NW Pakistan were produced using PSHA for an assumed
bedrock condition with V
s,30
= 760 m/s (NEHRP B/C boundary) on a 0.1 0.1 grid spacing.
Figure 5.16 through Figure 5.18 provide seismic hazard maps of PGA for 475-, 975-, and 2475-
year return periods, respectively. As higher ground accelerations are expected along the active
faults, the hazard map contours generally envelope known faults. As anticipated, the contours are
wider on the hanging wall side of reverse and normal faults because higher ground accelerations
are predicted on the hanging-wall side than on the footwall side. The largest PGAs (> 0.8g for a
475-year return period) are predicted along the MMT. The next highest PGAs are along the MBT
east segment (> 0.6 g) and the MBT west segment (> 0.4g), both for a 475-year return period.
The PGAs are computed to be relatively low near the southern faults, including the Kurram
Thrust and Salt Range Thrust because of their low activity rates.
Figure 5.19 through Figure 5.24 provide seismic hazard maps for SAs (T = 0.2 sec and
1.0 sec) for return periods of 475, 975, and 2475 years. These maps show trends similar to PGA,

59

with the largest SAs occurring along the MMT. Maximum SAs exceed 2.0g, 2.8g, and 4.0g at T
= 0.2 sec for 475-, 975-, and 2475-year return periods, respectively. Again, the hazard maps for
SA at T = 1.0 sec are similar to those for PGA, with the largest SAs values occurring along the
MMT.
5.8 Conditional mean spectra (CMS)
5.8.1 The CMS framework
The uniform hazard spectrum from a PSHA is often used to develop spectrum-compatible
ground motions for engineering analysis. This UHS is constructed by enveloping the spectral
amplitudes predicted by PSHA at all periods. The amplitude has a corresponding standard
normal random variable, or , computed as:

( )
( ) ( )
( ) T
T R M T Sa
T
Sa
w Sa
ln
ln
, , ln
o

c

= Eq. (5-1)

where ( ) T Sa ln is the natural logarithm of the spectral acceleration at the period of interest; and
( ) T R M
w Sa
, ,
ln
and ( ) T
Sa ln
o are the predicted mean and standard deviation, respectively, of
( ) T Sa ln

(McGuire 1995).
No single earthquake will produce a response spectrum as high as the UHS throughout
the frequency range considered (Baker and Cornell 2006). For example, the high- and low-
frequency portions of a UHS often correspond to different events (i.e., the high-frequency
portion of the UHS is often dominated by small nearby earthquakes, while the low-frequency
portion of the UHS is often dominated by larger, more distant earthquakes), Therefore, using a
UHS as a target or design spectrum, can be conservative if the structures at a site have a narrow
range of natural frequencies. A conditional mean spectrum (CMS) accounts for the variation of
SA amplitudes and for periods of interest, while maintaining the rigor of PSHA. The CMS
yields a spectrum that is smaller than the UHS, allowing a more efficient seismic design. The
procedure for computing CMS is summarized as follows (Baker 2008).

1) Determine the target SA and the associated M
w
, R, and at the period of interest (T*).
2) Compute the median and standard deviation of the response spectrum, given M
w
and R.

60

3) Compute at other periods, given (T*).
4) Compute the CMS.
5.8.2 Application of CMS to two cities in NW Pakistan
Step 1: Determine the target SA and the associated M
w
, R, and at T*. The UHS for the
cities of Islamabad and Muzaffarabad were developed using PSHA (Figure 5.1b and Figure
5.3b). For this example, periods of interest were selected as T* = 0.05 sec, 0.2 sec, and 1.0 sec.
Since the target SA(T*) was obtained from PSHA, the associated M
w
, R, and c values can be
taken as the mean values from deaggregation. These values are provided in Table 5.1 for return
periods of 475, 975, and 2475 years.
Step 2: Compute the median and standard deviation of the response spectrum, given M
w

and R. The Next Generation Attenuation (NGA) GMPEs used for the seismic hazard analysis
were employed in this step. These GMPEs are: Abrahamson-Silva (2008), Boore-Atkinson
(2008), Campbell-Bozorgnia (2008), and Idriss (2008). The CMS were developed separately
using each GMPE and later combined using equal weights.
Step 3: Compute at other periods, given (T*). Baker (2008) indicated that conditional
mean at other periods can be computed as:

( ) ( )
( ) ( )
* *
,
*
T T T
i
T T
i
c
c c
=
Eq. (5-2)


where
( ) ( )
*
T T
i
c c
is the mean value of ( )
i
T c , given ( )
*
T c ; and ( )
*
,T T
i
is the correlation
coefficient between the values at the two periods. Baker (2008) proposed the following relation
to define the correlation coefficient:

( )
( )
|
|
.
|

\
|
|
.
|

\
|
+ =
<
min
max min
189 . 0 max min
ln
189 . 0
ln 163 . 0 359 . 0
2
cos 1 ,
min
T
T T
I T T
T
t
Eq. (5-3)

where
min
T and
max
T are the smaller and larger of the two periods of interest, respectively, and
( ) 189 . 0
min
< T
I is a binary function equal to 1 if sec 189 . 0
min
< T and equal to 0 otherwise.

61

Equation (3) is valid only for the 0.05 < T (sec) < 5. Baker and Jayaram (2008) proposed
a refined correlation model that is valid over a wider period range of 0.01 to 10 seconds as:

if T
max
< 0.109 sec,
( ) ( ) 2 ,
2 1
C
T T
=
c c


else if T
min
> 0.109 sec,
( ) ( ) 1 ,
2 1
C
T T
=
c c


else
( ) ( ) 4 ,
2 1
C
T T
=
c c

Eq. (5-4)

where
( )
|
|
.
|

\
|
|
|
.
|

\
|
=
109 . 0 , max
ln 366 . 0
2
cos 1
min
max
1
T
T
C
t

Eq. (5-5)

<
|
|
.
|

\
|

|
.
|

\
|
+

=

otherwise
T if
T
T T
e
C
T
0
sec 2 . 0
0099 . 0
1
1
1 105 . 0 1
max
max
min max
100
2
5
max Eq. (5-6)

<
=
otherwise C
T if C
C
1
max 2
3
sec 109 . 0
Eq. (5-7)

( )
|
|
.
|

\
|
|
.
|

\
|
+ + =
109 . 0
cos 1 5 . 0
max
3 3 1 4
T
C C C C
t
Eq. (5-8)

where T
min
= min(T
1
,T
2
), and T
max
= max(T
1
,T
2
).
Figure 5.25 compares these two correlation coefficients.
Step 4: Compute CMS. Using the correlation coefficient proposed by Baker and Jayaram
(2008), the CMS at three target periods (T* = 0.05 sec, 0.2 sec, and 1.0 sec) corresponding to the
UHS for 2475-, 975-, and 475-year return periods for Islamabad and Muzaffarabad were
constructed as shown in Figure 5.26 and Figure 5.27, respectively.
For Islamabad, when the SA at T* = 0.05 sec with 475-year return period is selected, the
computed CMS is similar to the UHS. The SA for the CMS is only slightly smaller
(approximately 0.1 g) than the UHS SA at longer periods. When SA at T* = 0.2 sec with 475-

62

year return period is considered, the CMS becomes slightly smaller than the UHS at periods of
0.05 sec and 1.0 sec. However the CMS becomes similar to the UHS for PGA and SA at periods
> 3.0 sec. The CMS becomes significantly smaller than the UHS at T < 1.0 sec when the SA at
T* = 1.0 sec is targeted. The SA at T = 0.2 sec and the PGA are reduced by approximately 0.35g
and 0.13 g, respectively. A similar trend can be observed when the UHS for 975-year return
period is targeted at periods of 0.05 sec, 0.2 sec, and 1.0 sec, but the difference between the CMS
and the UHS becomes larger. When the CMS is generated for SA at T* = 1.0 sec, the SA at T =
0.2 sec and the PGA are reduced by approximately 0.46g and 0.27g, respectively. For the UHS
with 2475-year return period, the difference between the UHS and the CMS is substantial.
Especially when the SA at T* = 1.0 sec is targeted, the SA at T* = 0.2 sec and the PGA are
reduced to approximately 0.86g and 0.33g, respectively.
The CMS for Muzaffarabad show greater differences than that for Islamabad (Figure
5.27). When the SA at T* = 1.0 sec for 2475-year return period is targeted, the SA at T* = 0.2
sec and the PGA are reduced by a factor of approximately 2. When the SAs at T* = 0.05 sec and
0.2 sec were targeted, the SA at longer period is significantly reduced.
5.9 Summary and conclusion
Using the procedures described in Chapter 3 and Chapter 4, both probabilistic and
deterministic seismic hazard analyses performed for the 11 cities. The highest PGAs are
estimated to occur in Kaghan and Muzaffarabad, with PGAs for a 2475-year return period
exceeding 1.0g as a result of their proximity to major active faults. The cities of Astor, Bannu,
Malakand, Mangla, Peshawar, and Talagang are located much farther from significant faults;
therefore, PGAs for these cities are considerably lower. In general, median accelerations agree
well with those corresponding to 475- or 975-year return periods, while the 84
th
percentile
accelerations are comparable to those corresponding to the 2475-year return period.
PGAs estimated in this study using individual faults show distinct differences at cities
located at various source-to-site distances, in contrast to previous studies that used areal source
zones and computed PGAs that rarely exceeded 0.2 g and showed little variation among cities.
PGAs computed for cities located far from faults are similar to or slightly higher than those
computed using areal source zones, whereas PGAs computed for cities located close to faults are
2 to 4 times greater than those predicted by previous studies using areal source zones.

63

This chapter also presents the hazard maps. As higher ground accelerations are expected
along the active faults, the hazard map contours generally envelope known faults. The largest
PGAs (> 0.8g for a 475-year return period) are predicted along the MMT. The next highest
PGAs are observed along the MBT east segment (> 0.6 g) and the MBT west segment (> 0.4g),
both for a 475-year return period. It should be noted that the hazard maps are generated based on
the identified faults in NW Pakistan, and it is possible that the hazard is underestimated if there
are missing faults in the region.
The concept of CMS proposed by Baker (2008) was applied to the UHS for cities of
Islmabad and Muzaffarabad. The CMS at T* = 0.05 sec and 0.2 sec are not significantly different
from the UHS. However, it the difference between the UHS and the CMS is substantial when SA
at T* = 1.0 sec is targeted.


64

Table 5.1 Mean moment magnitude and distance from deaggregation analysis for cities of
Islamabad and Muzaffarabad. For Islamabad, all earthquakes correspond to the MBT fault, and
for Muzaffarabad, all earthquakes correspond to the Riasi fault.
Return
period
(years)
Period
(sec)
Islamabad Muzaffarabad
Mean
magnitude
(M
w
)
Mean
distance
(km)
Mean
epsilon
Mean
magnitude
(M
w
)
Mean
distance
(km)
Mean
epsilon
475
0.05 6.8 11 0.53 6.1 4.8 1.38
0.2 6.9 11 0.53 6.4 5.1 1.25
1.0 7.3 22 0.51 7.1 9.4 0.95
975
0.05 7.0 8.1 0.72 6.2 4.4 1.64
0.2 7.1 8.7 0.71 6.5 4.4 1.54
1.0 7.4 14 0.67 7.1 7.6 1.22
2475
0.05 7.2 6.5 1.09 6.3 4.0 1.96
0.2 7.2 6.7 1.07 6.5 3.9 1.86
1.0 7.5 8.8 0.98 7.1 6.1 1.55


65


Figure 5.1 Seismic hazard analysis results for Islamabad for an assumed bedrock with V
s,30
of
760 m/s. (a) PSHA hazard curves for PGA at 5% damping. The hazards from the top five
contributing faults are shown; (b) PSHA uniform hazard spectra (UHS) at various return periods
and DSHA hazard spectra.
0 0.5 1 1.5
Peak ground acceleration (g)
0.0001
0.001
0.01
0.1
1
A
n
n
u
a
l

f
r
e
q
u
e
n
c
y

o
f

e
x
c
e
e
d
a
n
c
e
Total Hazard
Jhelum Fault
MBT (west)
MMT
Puran Fault
Riwat Thrust
R
e
t
u
r
n

p
e
r
i
o
d

(
y
e
a
r
)
475
975
2475
0.01 0.1 1 10
Period (sec)
0
1
2
3
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
PSHA (2475 years)
PSHA (975 years)
PSHA (475 years)
DSHA (84 percentile)
DSHA (Median)
(a) PSHA hazard curve
(b) Spectra

66


Figure 5.2 Deaggregation of seismic hazard for Islamabad for 475-year return period at three spectral periods: (a) 0.05 sec, (b) 0.2 sec,
and (c) 1.0 sec.

67


Figure 5.3 Seismic hazard analysis results for Muzaffarabad for an assumed bedrock with V
s,30
of
760 m/s. (a) PSHA hazard curves for PGA at 5% damping. The hazards from the top five
contributing faults are shown; (b) PSHA uniform hazard spectra (UHS) at various return periods
and DSHA hazard spectra.
0 0.5 1 1.5
Peak ground acceleration (g)
0.0001
0.001
0.01
0.1
1
A
n
n
u
a
l

f
r
e
q
u
e
n
c
y

o
f

e
x
c
e
e
d
a
n
c
e
Total Hazard
Jhelum Fault
Mansehra Thrust
MMT
Puran Fault
Riasi Thrust
R
e
t
u
r
n

p
e
r
i
o
d

(
y
e
a
r
)
475
975
2475
0.01 0.1 1 10
Period (sec)
0
1
2
3
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
PSHA (2475 years)
PSHA (975 years)
PSHA (475 years)
DSHA (84 percentile)
DSHA (Median)
(a) PSHA hazard curve
(b) Spectra

68


Figure 5.4 Deaggregation of seismic hazard for Muzaffarabad for 475-year return period at three spectral periods: (a) 0.05 sec, (b) 0.2
sec, and (c) 1.0 sec.


69


Figure 5.5 Seismic hazard analysis results for Kaghan for an assumed bedrock with V
s,30
of 760
m/s. (a) PSHA hazard curves for PGA at 5% damping. The hazards from the top five contributing
faults are shown; (b) PSHA uniform hazard spectra (UHS) at various return periods and DSHA
hazard spectra.
0 0.5 1 1.5
Peak ground acceleration (g)
0.0001
0.001
0.01
0.1
1
A
n
n
u
a
l

f
r
e
q
u
e
n
c
y

o
f

e
x
c
e
e
d
a
n
c
e
Total Hazard
Balakot Shear Zone
Mansehra Thrust
MMT
MCT
Puran Fault
R
e
t
u
r
n

p
e
r
i
o
d

(
y
e
a
r
)
475
975
2475
0.01 0.1 1 10
Period (sec)
0
1
2
3
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
PSHA (2475 years)
PSHA (975 years)
PSHA (475 years)
DSHA (84 percentile)
DSHA (Median)
(a) PSHA hazard curve
(b) Spectra

70


Figure 5.6 Seismic hazard result for Peshawar for an assumed bedrock with V
s,30
of 760 m/s. (a)
PSHA hazard curves for PGA at 5% damping. The hazards from the top five contributing faults
are shown; (b) PSHA uniform hazard spectra (UHS) at various return periods and DSHA hazard
spectra.
0 0.5 1 1.5
Peak ground acceleration (g)
0.0001
0.001
0.01
0.1
1
A
n
n
u
a
l

f
r
e
q
u
e
n
c
y

o
f

e
x
c
e
e
d
a
n
c
e
Total Hazard
Hissartang Fault
MBT (west)
MMT
Nowshera Fault
Puran Fault
R
e
t
u
r
n

p
e
r
i
o
d

(
y
e
a
r
)
475
975
2475
0.01 0.1 1 10
Period (sec)
0
1
2
3
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
PSHA (2475 years)
PSHA (975 years)
PSHA (475 years)
DSHA (84 percentile)
DSHA (Median)
(a) PSHA hazard curve
(b) Spectra

71


Figure 5.7 Comparisons of spectral accelerations computed by PSHA at 475-, 975-, and 2475-
year return periods with median and 84
th
percentile spectral accelerations computed by DSHA
for 11 cities in NW Pakistan for an assumed bedrock with V
s,30
of 760 m/s. (a) PGA; (b) SA (T =
0.2 sec); and (c) SA (T = 1.0 sec)
0
0.5
1
1.5
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
0
0.5
1
1.5
2
2.5
3
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
0
0.5
1
1.5
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
A
s
t
o
r
B
a
l
a
k
o
t
B
a
n
n
u
I
s
l
a
m
a
b
a
d
K
a
g
h
a
n
K
o
h
a
t
M
a
l
a
k
a
n
d
M
a
n
g
l
a
M
u
z
a
f
f
a
r
a
b
a
d
P
e
s
h
a
w
a
r
T
a
l
a
g
a
n
g
PSHA
475-yr return period
975-yr return period
2475-yr return period
DSHA
84 percentile
Median
(a) PGA
(b) SA (T=0.2 sec)
(c) SA (T=1.0 sec)

72


Figure 5.8 Seismic hazard result for Astor for an assumed bedrock with V
s,30
of 760 m/s. (a)
PSHA hazard curves for PGA at 5% damping. The hazards from the top five contributing faults
are shown; (b) PSHA uniform hazard spectra (UHS) at various return periods and DSHA hazard
spectra.
0 0.5 1 1.5
Peak ground acceleration (g)
0.0001
0.001
0.01
0.1
1
A
n
n
u
a
l

f
r
e
q
u
e
n
c
y

o
f

e
x
c
e
e
d
a
n
c
e
Total Hazard
Jhelum fault
MCT
MMT
Raikot fault
Stak fault
R
e
t
u
r
n

p
e
r
i
o
d

(
y
e
a
r
)
475
975
2475
0.01 0.1 1 10
Period (sec)
0
1
2
3
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
PSHA (2475 years)
PSHA (975 years)
PSHA (475 years)
DSHA (84 percentile)
DSHA (Median)
(a) PSHA hazard curve
(b) Spectra

73


Figure 5.9 Seismic hazard result for Balakot for an assumed bedrock with V
s,30
of 760 m/s. (a)
PSHA hazard curves for PGA at 5% damping. The hazards from the top five contributing faults
are shown; (b) PSHA uniform hazard spectra (UHS) at various return periods and DSHA hazard
spectra.
0 0.5 1 1.5
Peak ground acceleration (g)
0.0001
0.001
0.01
0.1
1
A
n
n
u
a
l

f
r
e
q
u
e
n
c
y

o
f

e
x
c
e
e
d
a
n
c
e
Total Hazard
Balakot shear zone
Jhelum fault
MMT
Mansehra fault
Puran fault
R
e
t
u
r
n

p
e
r
i
o
d

(
y
e
a
r
)
475
975
2475
0.01 0.1 1 10
Period (sec)
0
1
2
3
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
PSHA (2475 years)
PSHA (975 years)
PSHA (475 years)
DSHA (84 percentile)
DSHA (Median)
(a) PSHA hazard curve
(b) Spectra

74


Figure 5.10 Seismic hazard result for Bannu for an assumed bedrock with V
s,30
of 760 m/s. (a)
PSHA hazard curves for PGA at 5% damping. The hazards from the top five contributing faults
are shown; (b) PSHA uniform hazard spectra (UHS) at various return periods and DSHA hazard
spectra.
0 0.5 1 1.5
Peak ground acceleration (g)
0.0001
0.001
0.01
0.1
1
A
n
n
u
a
l

f
r
e
q
u
e
n
c
y

o
f

e
x
c
e
e
d
a
n
c
e
Total Hazard
Karack fault
Khair-I-Murat fault
Kurram thrust
MMT
Surghar range thrust
R
e
t
u
r
n

p
e
r
i
o
d

(
y
e
a
r
)
475
975
2475
0.01 0.1 1 10
Period (sec)
0
1
2
3
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
PSHA (2475 years)
PSHA (975 years)
PSHA (475 years)
DSHA (84 percentile)
DSHA (Median)
(a) PSHA hazard curve
(b) Spectra

75


Figure 5.11 Seismic hazard result for Kohat for an assumed bedrock with V
s,30
of 760 m/s. (a)
PSHA hazard curves for PGA at 5% damping. The hazards from the top five contributing faults
are shown; (b) PSHA uniform hazard spectra (UHS) at various return periods and DSHA hazard
spectra.
0 0.5 1 1.5
Peak ground acceleration (g)
0.0001
0.001
0.01
0.1
1
A
n
n
u
a
l

f
r
e
q
u
e
n
c
y

o
f

e
x
c
e
e
d
a
n
c
e
Total Hazard
Hissartang fault
Khair-I-Murat fault
Khairrabad fault
MBT (west)
MMT
R
e
t
u
r
n

p
e
r
i
o
d

(
y
e
a
r
)
475
975
2475
0.01 0.1 1 10
Period (sec)
0
1
2
3
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
PSHA (2475 years)
PSHA (975 years)
PSHA (475 years)
DSHA (84 percentile)
DSHA (Median)
(a) PSHA hazard curve
(b) Spectra

76


Figure 5.12 Seismic hazard result for Malakand for an assumed bedrock with V
s,30
of 760 m/s. (a)
PSHA hazard curves for PGA at 5% damping. The hazards from the top five contributing faults
are shown; (b) PSHA uniform hazard spectra (UHS) at various return periods and DSHA hazard
spectra.
0 0.5 1 1.5
Peak ground acceleration (g)
0.0001
0.001
0.01
0.1
1
A
n
n
u
a
l

f
r
e
q
u
e
n
c
y

o
f

e
x
c
e
e
d
a
n
c
e
Total Hazard
MBT (west)
MMT
Nowshera fault
Mansehra thrust
Puran fault
R
e
t
u
r
n

p
e
r
i
o
d

(
y
e
a
r
)
475
975
2475
0.01 0.1 1 10
Period (sec)
0
1
2
3
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
PSHA (2475 years)
PSHA (975 years)
PSHA (475 years)
DSHA (84 percentile)
DSHA (Median)
(a) PSHA hazard curve
(b) Spectra

77


Figure 5.13 Seismic hazard result for Mangla for an assumed bedrock with V
s,30
of 760 m/s. (a)
PSHA hazard curves for PGA at 5% damping. The hazards from the top five contributing faults
are shown; (b) PSHA uniform hazard spectra (UHS) at various return periods and DSHA hazard
spectra.
0 0.5 1 1.5
Peak ground acceleration (g)
0.0001
0.001
0.01
0.1
1
A
n
n
u
a
l

f
r
e
q
u
e
n
c
y

o
f

e
x
c
e
e
d
a
n
c
e
Total Hazard
Dijabba fault
Jhelum fault
MBT (east)
Riasi thrust
Riwat thrust
R
e
t
u
r
n

p
e
r
i
o
d

(
y
e
a
r
)
475
975
2475
0.01 0.1 1 10
Period (sec)
0
1
2
3
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
PSHA (2475 years)
PSHA (975 years)
PSHA (475 years)
DSHA (84 percentile)
DSHA (Median)
(a) PSHA hazard curve
(b) Spectra

78


Figure 5.14 Seismic hazard result for Talagang for an assumed bedrock with V
s,30
of 760 m/s. (a)
PSHA hazard curves for PGA at 5% damping. The hazards from the top five contributing faults
are shown; (b) PSHA uniform hazard spectra (UHS) at various return periods and DSHA hazard
spectra.

0 0.5 1 1.5
Peak ground acceleration (g)
0.0001
0.001
0.01
0.1
1
A
n
n
u
a
l

f
r
e
q
u
e
n
c
y

o
f

e
x
c
e
e
d
a
n
c
e
Total Hazard
Dijabba fault
Jhelum fault
Khair-I-Murat fault
MBT (west)
Salt range thrust
R
e
t
u
r
n

p
e
r
i
o
d

(
y
e
a
r
)
475
975
2475
0.01 0.1 1 10
Period (sec)
0
1
2
3
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
PSHA (2475 years)
PSHA (975 years)
PSHA (475 years)
DSHA (84 percentile)
DSHA (Median)
(a) PSHA hazard curve
(b) Spectra

79


Figure 5.15 PGAs computed by PSHA and DSHA for an assumed bedrock with V
s,30
of 760 m/s,
compared to those computed by others (PMD and NORSAR 2006; Monalisa et al. 2007). PGAs
estimated by Monalisa et al. (2007) correspond to a 475-year return period, while PMD and
NORSAR (2006) values correspond to a 500-year return period.


0
0.2
0.4
0.6
0.8
P
e
a
k

g
r
o
u
n
d

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
A
s
t
o
r
B
a
n
n
u
I
s
l
a
m
a
b
a
d
K
a
g
h
a
n
K
o
h
a
t
M
a
l
a
k
a
n
d
M
a
n
g
l
a
M
u
z
a
f
f
a
r
a
b
a
d
P
e
s
h
a
w
a
r
T
a
l
a
g
a
n
g
Monalisa et al. (2007) (475-yr return period)
PMD and NORSAR (2006) (500-yr return period)
This study-PSHA (475-yr return period)
This study-DSHA (median)

80


Figure 5.16 Seismic hazard map of NW Pakistan for PGA for a 475-year return period. A 0.1
0.1 grid spacing was analyzed using PSHA for an assumed bedrock with V
s,30
of 760 m/s. The
contour interval is 0.2g.
Astor
Kaghan
Malakand
Balakot
Muzaffarabad
Islamabad
Kohat
Bannu
Talagang
Peshawar
Mangla
36 N
35 N
34 N
33 N
32 N
70 E 71 E 72 E 73 E 74 E 75 E
0.4
0.4
0.8
Reverse fault
Normal fault
Strike-slip fault
PGA (g)
1.2
1.0
0.8
0.6
0.4
0.2

81


Figure 5.17 Seismic hazard map of NW Pakistan for PGA for a 975-year return period. A 0.1
0.1 grid spacing was analyzed using PSHA for an assumed bedrock with V
s,30
of 760 m/s. The
contour interval is 0.2g.

Astor
Kaghan
Malakand
Balakot
Muzaffarabad
Islamabad
Kohat
Bannu
Talagang
Peshawar
Mangla
36 N
35 N
34 N
33 N
32 N
70 E 71 E 72 E 73 E 74 E 75 E
0.4
0
.
6
1.0
Reverse fault
Normal fault
Strike-slip fault
PGA (g)
1.2
1.0
0.8
0.6
0.4
0.2

82


Figure 5.18 Seismic hazard map of NW Pakistan for PGA for a 2475-year return period. A 0.1
0.1 grid spacing was analyzed using PSHA for an assumed bedrock with V
s,30
of 760 m/s. The
contour interval is 0.2g.
Astor
Kaghan
Malakand
Balakot
Muzaffarabad
Islamabad
Kohat
Bannu
Talagang
Peshawar
Mangla
36 N
35 N
34 N
33 N
32 N
70 E 71 E 72 E 73 E 74 E 75 E
Reverse fault
Normal fault
Strike-slip fault
PGA (g)
1.2
1.0
0.8
0.6
0.4
0.2

83


Figure 5.19 Seismic hazard map of NW Pakistan for SA (T=0.2 sec) for a 475-year return period.
A 0.1 0.1 grid spacing was analyzed using PSHA for an assumed bedrock with V
s,30
of 760
m/s. The contour interval is 0.4g.
Astor
Kaghan
Malakand
Balakot
Muzaffarabad
Islamabad
Kohat
Bannu
Talagang
Peshawar
Mangla
36 N
35 N
34 N
33 N
32 N
70 E 71 E 72 E 73 E 74 E 75 E
SA (g)
4.0
2.0
1.6
1.2
0.8
0.4
3.6
3.2
2.8
2.4
Reverse fault
Normal fault
Strike-slip fault

84


Figure 5.20 Seismic hazard map of NW Pakistan for SA (T=0.2 sec) for a 975-year return period.
A 0.1 0.1 grid spacing was analyzed using PSHA for an assumed bedrock with V
s,30
of 760
m/s. The contour interval is 0.4g.

Astor
Kaghan
Malakand
Balakot
Muzaffarabad
Islamabad
Kohat
Bannu
Talagang
Peshawar
Mangla
36 N
35 N
34 N
33 N
32 N
70 E 71 E 72 E 73 E 74 E 75 E
0.8
1
.
2
SA (g)
4.0
2.0
1.6
1.2
0.8
0.4
3.6
3.2
2.8
2.4
Reverse fault
Normal fault
Strike-slip fault

85


Figure 5.21 Seismic hazard map of NW Pakistan for SA (T=0.2 sec) for a 2475-year return
period. A 0.1 0.1 grid spacing was analyzed using PSHA for an assumed bedrock with V
s,30

of 760 m/s. The contour interval is 0.4g.
Astor
Kaghan
Malakand
Balakot
Muzaffarabad
Islamabad
Kohat
Bannu
Talagang
Peshawar
Mangla
36 N
35 N
34 N
33 N
32 N
70 E 71 E 72 E 73 E 74 E 75 E
0.8
1
.
6
SA (g)
4.0
2.0
1.6
1.2
0.8
0.4
3.6
3.2
2.8
2.4
Reverse fault
Normal fault
Strike-slip fault

86


Figure 5.22 Seismic hazard map of NW Pakistan for SA (T=1.0 sec) for a 475-year return period.
A 0.1 0.1 grid spacing was analyzed using PSHA for an assumed bedrock with V
s,30
of 760
m/s. The contour interval is 0.2g.

Astor
Kaghan
Malakand
Balakot
Muzaffarabad
Islamabad
Kohat
Bannu
Talagang
Peshawar
Mangla
36 N
35 N
34 N
33 N
32 N
70 E 71 E 72 E 73 E 74 E 75 E
0.4
Reverse fault
Normal fault
Strike-slip fault
PGA (g)
1.2
1.0
0.8
0.6
0.4
0.2

87


Figure 5.23 Seismic hazard map of NW Pakistan for SA (T=1.0 sec) for a 975-year return period.
A 0.1 0.1 grid spacing was analyzed using PSHA for an assumed bedrock with V
s,30
of 760
m/s. The contour interval is 0.2g.

Astor
Kaghan
Malakand
Balakot
Muzaffarabad
Islamabad
Kohat
Bannu
Talagang
Peshawar
Mangla
36 N
35 N
34 N
33 N
32 N
70 E 71 E 72 E 73 E 74 E 75 E
0.4
0.4
0.6
0.6
Reverse fault
Normal fault
Strike-slip fault
PGA (g)
1.2
1.0
0.8
0.6
0.4
0.2

88


Figure 5.24 Seismic hazard map of NW Pakistan for SA (T=1.0 sec) for a 2475-year return
period. A 0.1 0.1 grid spacing was analyzed using PSHA for an assumed bedrock with V
s,30

of 760 m/s. The contour interval is 0.2g.



Astor
Kaghan
Malakand
Balakot
Muzaffarabad
Islamabad
Kohat
Bannu
Talagang
Peshawar
Mangla
36 N
35 N
34 N
33 N
32 N
70 E 71 E 72 E 73 E 74 E 75 E
0.6
0
.
8
Reverse fault
Normal fault
Strike-slip fault
PGA (g)
1.2
1.0
0.8
0.6
0.4
0.2

89


Figure 5.25 Comparison of correlation coefficient proposed by Baker (2008) and Baker and
Jayaram (2008).
0.01 0.1 1 10
Period (sec)
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
C
o
r
r
e
l
a
t
i
o
n

c
o
e
f
f
i
c
i
e
n
t
,

Baker and Jayaram


(2008)
T*=0.05 sec
T*=0.1 sec
T*=0.2 sec
T*=0.5 sec
T*=2.0 sec
T*=10 sec
Baker (2008)
T*=0.1 sec
T*=0.2 sec
T*=0.5 sec
T*=2.0 sec

90


Figure 5.26 CMS at T* = 0.05, 0.2, and 1 second for 475-, 975-, and 2475-year return periods in
Islamabad.
0
0.4
0.8
1.2
1.6
2
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
UHS (475 years)
CMS
0
0.4
0.8
1.2
1.6
2
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
UHS (975 years)
CMS
0.01 0.1 1 10
Period (sec)
0
0.4
0.8
1.2
1.6
2
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
UHS (2475 years)
CMS

91


Figure 5.27 CMS at T* = 0.05, 0.2, and 1 second for 475-, 975-, and 2475-year return periods in
Muzaffarabad.
0
0.4
0.8
1.2
1.6
2
2.4
2.8
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
UHS (475 years)
CMS
0
0.4
0.8
1.2
1.6
2
2.4
2.8
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
UHS (975 years)
CMS
0.01 0.1 1 10
Period (sec)
0
0.4
0.8
1.2
1.6
2
2.4
2.8
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
UHS (2475 years)
CMS

92

CHAPTER 6. CURRENT SEISMIC SLOPE STABILITY APPROACH
6.1 Introduction
This chapter introduces the characteristics of earthquake-induced landslides and general
approaches to analyze seismic slope stability (pseudo-static analysis, the Newmark (1965)
sliding block method, modified Newmark method, simplified displacement charts, and energy-
based methods). This study focuses on whether slopes fail during earthquakes rather than
estimating displacements. Therefore, the pseudo-static slope stability analysis is selected to
evaluate the regional performance of slopes during an earthquake, and discussed in detail.
Although sophisticated analyses (i.e., 2D and 3D limit equilibrium, 2D and 3D finite element,
and 2D and 3D discrete element analyses) are available to analyze the stability of individual
slopes, these are not practical to assess the regional performance of slopes. Therefore, this study
presents a simplified seismic slope stability analysis using estimated rock properties and slope
geometries in the affected areas. To better understand landslide distributions and help explain the
high concentration of landslides in the near-field, this study proposes several factors that are not
commonly incorporated in current slope stability practice.
6.2 Earthquake-induced landslides
Earthquake-induced landslides can cause numerous fatalities and can create massive
disruption following an earthquake as a result of blocking critical transportation routes in
mountainous terrain, damming waterways, and triggering seiches or tsunamis. For example, the
2008 Wenchuan earthquake (moment magnitude, M
w
7.9) in China triggered more than 15,000
landslides, rockfalls, and debris flows, resulting in approximately 20,000 deaths (Yin et al. 2009).
The 2005 Kashmir, Pakistan, earthquake (M
w
7.6) triggered more than 2,400 landslides (Sato et
al. 2007), including the 6810
6
m
3
Hattian Bala rock avalanche which destroyed an entire village
and caused about 1,000 fatalities (Dunning et al. 2007). The 1999 Chi-chi, Taiwan, earthquake
(M
w
7.6) triggered more than 9,200 landslides throughout a region of approximately 128 km
2

(Liao 2000). The largest of these was the Tsaoling landslide which involved 12510
6
m
3
of rock
and caused 39 fatalities (Tang et al. 2009). The 1989 Loma Prieta, California, earthquake (M
w

6.9) triggered thousands of landslides over an area of about 15,000 km
2
, damaging more than
200 residences, numerous roads, and other structures, with damage estimates exceeding US$30
million (Keefer and Manson 1998).

93

Material type (soil or rock) and the character of movement are used to classify
seismically-induced landslides as shown in Table 6.1 and Table 6.2 (Keefer 2002). For example,
disrupted landslides involve broken, sheared, and disturbed material, while coherent
landslides involve a few relatively intact blocks that rotate and/or translate on a failure surface,
and generally exhibit lower velocities than disrupted slides (Keefer 1984). Numerous factors
influence the seismic stability of slopes and the landslide characteristics, including the geologic
and hydrologic conditions (e.g., weathering); topography: climate; and the strength of shaking
(i.e., ground motion amplitude and frequency content; earthquake magnitude). Factors such as
topsoil thickness and vegetation have no significant impact on the earthquake induced landslides,
although they become more important for rainfall triggered landslides (Khazai and Sitar 2004).
Because earthquake-induced landslides cause significant damage, the distribution of
landslide has been of interest for many scientists and engineers. One of the pioneering researches
was conducted by Keefer (1984). He proposed the relation between earthquake magnitude and
area affected by landslides based on historical earthquake data. According to his relation, the
area of approximately 100,000 km
2
can be possibly affected by landslides induced by the
earthquake of M
w
= 8.0 (Figure 6.1). He also proposed the relation between maximum distance
of landslides from fault and earthquake magnitude as shown in Figure 6.2. Keefer (1984)
illustrated the minimum earthquake magnitudes required to trigger different types of landslides
as shown in Table 6.3. Rock falls, rock slides, soil falls, and disrupted soil slides can be triggered
by small magnitude earthquake (M
w
= 3.6), whereas rock avalanches and soil avalanches require
large magnitude (M
w
= 5.7 and 6.3, respectively). Similar correlations have related landslide
concentration to distance from fault (Keefer 2000; Khazai and Sitar 2004; Sato et al. 2007; Dai et
al. 2011). These studies suggest that landslides are highly concentrated near the fault (near-field).
6.3 General approaches to analyze seismic slope stability
Several simplified methods are available to estimate the factor of safety (FS) against
sliding and displacements during earthquake-induced landslides. These methods include pseudo-
static analysis, the Newmark (1965) sliding block method, modified Newmark methods,
simplified displacement charts, and energy-based methods.
Terzaghi (1950) first applied a pseudo-static approach to analyze seismic slope stability.
This approach is based on conventional limit equilibrium slope stability analysis with horizontal

94

and vertical pseudo-static forces that act on the critical failure mass. This approach predicts the
FS against sliding, and can be also used to determine a yield acceleration (a
y
) for use in a
Newmark sliding block analysis to estimate displacements.
Newmark (1965) developed a sliding block analysis to estimate permanent displacements
of a sliding mass. In this method, the sliding mass is represented as a rigid block on an inclined
plane and displacements are integrated from base acceleration pulses that exceed the blocks
yield acceleration. If the acceleration does not exceed the yield acceleration, there is no
computed sliding block displacement.
Makdisi and Seed (1978) proposed a simple method for the seismic design of
embankments, that uses the concept originally proposed by Newmark (1965). They used a finite
element analysis to compute the variation of permanent displacement with a
y
/a
max
(where a
max
is
the peak ground acceleration) and earthquake magnitude. The analyses employed several real
and hypothetical dams with heights ranging from 30 to 60 m and constructed of compacted fine-
grained soils or very dense coarse-grained soils that were subjected to several recorded and
synthetic ground motions scaled to represent different earthquake magnitudes. The displacement
was normalized by the peak base acceleration and fundamental period of the embankment.
Yegian et al. (1991) and Bray et al. (1998) proposed variations of this method to estimate
seismic slope displacement.
When seismic waves propagate through a slope, different parts of the slope will move by
different amounts and with different phases (Kramer and Smith 1997). This phenomenon is most
significant for thick, soft materials and short wavelengths. To accommodate these factors, many
researchers have modified the Newmark method to accommodate compliant slope materials
(e.g., Lin and Whitman 1983; Kramer and Smith 1997; Rathje and Bray 2000). In addition,
Matasovic et al. (1997) incorporated a degrading yield acceleration into the Newmark method to
accommodate potential cyclic degradation of soil shear strength.
Kokusho and Kabasawa (2003) proposed an energy-based approach, and Kokusho and
Ishizawa (2007) further developed it. Specifically, their model accounts for potential energy of
the sliding mass, earthquake energy contributing to slope failure, dissipated energy due to sliding
mass deformation, and kinetic energy of the sliding mass. They correlated these energies with
residual horizontal displacement, material properties (mass of soil block and friction angle), and

95

average slope angle. Consequently, they found that the residual horizontal displacement can be
estimated from the earthquake energy.
This study focuses on whether slopes fail during earthquakes rather than estimating
displacements. Therefore, the factor of safety against sliding using a pseudo-static slope stability
analysis approach is considered. Furthermore, this approach is applied regionally to evaluate
landslide concentration, and the analyses described below were not derived from any particular
slope.
6.4 Pseudo-static slope stability analysis
As noted above, Terzaghi (1950) first applied a pseudo-static approach to analyze seismic
slope stability. This approach uses a single, monotonically-applied horizontal and/or vertical
acceleration to represent earthquake loading. (Although the vertical acceleration can be included
in a pseudo-static analysis, it is rarely used in practice, as explained below.) The horizontal and
vertical pseudo-static forces, F
h
and F
v
, respectively, act through the sliding mass centroid and
are defined as:

W k
g
W a
F and W k
g
W a
F
v
v
v h
h
h
= = = =

Eq. (6-1)

where a
h
and a
v
= horizontal and vertical accelerations, respectively; k
h
and k
v
= dimensionless
horizontal and vertical pseudo-static coefficients, respectively; and W = weight of the failure
mass. Using an infinite slope analysis with this approach has the benefits of being simple and
able to reasonably approximate shallow slope failures, including earthquake-induced shallow,
disrupted landslides which constitute most of the landslides in the study areas. Due to the small
thickness of landslide mass, it was assumed that the water table was below the sliding plane. For
an infinite slope where horizontal and vertical pseudo-static seismic loads act through the sliding
mass centroid, FS is calculated as:

( ) | |
( ) i k i k
D
c
i k i k
FS
h v
t
h v
cos sin 1
tan sin cos 1
+
'
+ '
=

|

Eq. (6-2)


96

where c and | are Mohr-Coulomb strength parameters that describe the shear strength on the
failure plane; = unit weight of failure mass; i = slope and failure plane angles; and D = failure
mass thickness.
The horizontal pseudo-static force has a larger influence on the FS than the vertical
pseudo-static force, as F
h
reduces the resisting force and increases the driving force. Thus,
selecting appropriate values for k
h
is important for estimating meaningful values of FS. Many
investigators have suggested values of pseudo-static coefficients. For example, Terzaghi (1950)
proposed k
h
= 0.1 for severe earthquakes (Rossi-Forel intensity IX); k
h
= 0.2 for violent,
destructive earthquakes (Rossi-Forel X); and k
h
= 0.5 for catastrophic earthquakes. Seed
(1979) suggested k
h
= 0.10 (M = 6.5) to 0.15 (M = 8.25) for earth dams constructed of ductile
soils with crest accelerations less than 0.75g. Later, Pyke (1991) proposed a relation between
pseudo-static coefficient and M
w
, where k
h
approaches 0.5 for M
w
> 8.0.
Other researchers have proposed pseudo-static coefficients that vary with anticipated
maximum horizontal acceleration, a
h,max
at the base of sliding mass. For example, Marcuson and
Curro (1981) suggested k
h
= (1/3 to 1/2)(a
h,max
/g). Hynes-Griffin and Franklin (1984) suggested
k
h
= 0.5(a
h,max
/g). Nozu et al. (1997) proposed a slightly different form [k
h
=1/3 (a
h,max
/g)
1/3
],
which gives higher pseudo-static coefficient than the former two approaches for a
h,max
0.5g,
and falls between the other two approaches for a
h,max
> 0.5g. Anderson et al. (2008) pointed out
that the pseudo-static coefficient is typically assumed to be less than 50 percent of a
h,max
, but is a
function of the slope height and frequency content of the ground motion. Two conventionally-
used horizontal pseudo-static coefficients, k
h
= 0.5(a
h,max
/g) and k
h
= 0.33(a
h,max
/g) were
employed. In addition, k
h
= 0.9(a
h,max
/g) was considered to account for shallow slides where
wave incoherency and scattering effects may be less prevalent than in large landslides. Equal
weights were assigned to these three values of k
h
for statistical analysis.
Many researchers have discounted vertical acceleration in computing FS because, as
noted by Kramer (1996), the vertical pseudo-static force reduces both driving and resisting forces.
The effect of discounting the vertical pseudo-static force is examined subsequently.

97

6.5 Evaluating the pseudo-static approach using landslides triggered by the 2005
Kashmir, Pakistan, earthquake
The 2005 Kashmir earthquake (M
w
=7.6) occurred in the northern Pakistan where thrust
movements dominant due to the Indian plate colliding with the Eurasian plate. The Balakot-
Garhi fault which ruptured during the earthquake is a 50-km-long thrust fault with a dip of
30NE and rupture depth of 26 km. Sato et al. (2007) reported that 2,424 landslides occurred in
their 55 km x 51 km study area. As shown in Figure 6.3, the landslides are highly concentrated
along the fault. Sato et al. (2007) performed a statistical analysis of the distribution of landslides
triggered by the earthquake. Based on these data, the landslide distribution was expressed as a
landslide concentration (LC), which is defined as the number of landslides per square kilometer
of surface area (Keefer 2000). The rock slopes in the area affected by the earthquake are mainly
sedimentary, igneous, and metamorphic rocks of Precambrian to Quaternary ages (Sato et al.
2007). Landslides occurred most frequently in the Tertiary-age Murree Formation (Tmm) which
consists of hard sandstone and siltstone with thin intraformational conglomerate lenses (Sato et
al. 2007). Kamp et al. (2008) described the failed sedimentary rocks as undeformed to tightly-
folded, highly-cleaved and fractured.
To perform a pseudo-static slope stability analysis, a number of rock properties are
needed, as well as geometric properties of the slope. For individual slopes, these values are
evaluated as part of a site characterization program. However, for this study, the regional
performance of rock slopes is of interest; therefore, this study characterizes these parameters in
terms of typical mechanical and geometric properties that apply to the affected region as a whole.
The rock properties needed for the analysis are total unit weight, , and Mohr-Coulomb shear
strength parameters, |' and c'. The geometric parameters needed are D
t
and i (defined above).
Because Mohr-Coulomb strength properties for rock specimens are seldom measured directly
(other than for joints), the Hoek-Brown strength criterion (Hoek et al. 2002) was used to estimate
appropriate Mohr-Coulomb strength parameters. The Hoek-Brown criterion requires the
following parameters (Hoek et al. 2002): rock uniaxial unconfined compressive strength, o
c
;
material constant for intact rock, m
i
; Geological Strength Index, GSI; and disturbance factor, DF,
that accounts for the degree of disturbance to the rock mass from blasting and/or stress relief.
These parameters were estimated from reported geologic and geotechnical conditions in the
affected region. The Hoek-Brown strength parameters were then converted to equivalent Mohr-

98

Coulomb shear strength parameters over the appropriate effective confining stress range for use
in the pseudo-static slope stability analysis.
As discussed by Hoek et al. (2002), the degree of weathering plays an important role in
estimating rock strength parameters. Sato et al. (2007) reported that 79% of the landslides were
shallow, disrupted rockslides, and Owen et al. (2008) observed that these shallow landslides
typically involved the top few meters of weathered bedrock, regolith, and soil. Figure 6.4 shows
the picture of shallow landslides triggered by 2005 Kashmir earthquake. For the Mohr-Coulomb
strength parameters, it was assumed that the values computed using the Hoek-Brown criterion
were mean values, and standard deviations were estimated using coefficients of variation (COV)
of 5%, 30%, and 15% for , c', and |', respectively, as suggested by Lee et al. (1983). Using the
Hoek-Brown parameters in Table 6.4 for the Kashmir earthquake case yield |' = 40 6 and c' =
27 8 kPa. Detailed explanation is presented in Appendix A.
Based on the observation by Owen et al. (2008), D
t
= 1 to 5 m were considered. Figure
6.5 (a) shows the distribution of slope angles for slope that failed during the 2005 Kashmir
earthquake (Kamp et al. 2008). Dai and Wang (1992) suggested that 99.73% of all values of a
normally distributed parameter fall within three standard deviations of the mean. Assuming a
mean value of 3 m and minimum and maximum values of 1 and 5 m, respectively, the three-
sigma rule was applied to estimate a standard deviation of 0.67 m. Figure 6.5(a) shows the
distribution of slope angles for slopes that failed during the 2005 Kashmir earthquake (Kamp et
al. 2008). This study considered a normal distribution with a mean slope angle of 28 and a
standard deviation of 9 to represent the failed slope data. The actual cumulative distribution of
documented slope angles is similar to the theoretical cumulative normal distribution as shown in
Figure 6.5(a). Table 6.4 summarizes the geometric parameters for this case.
Figure 6.6 shows the horizontal ground accelerations predicted using four GMPEs
released by Next Generation Attenuation (NGA) projects (Abrahamson and Silva 2008; Boore
and Atkinson 2008; Campbell and Bozorgnia 2008; Idriss 2008). Figure 6.7 compares the static
and pseudo-static FS against sliding, using mean material properties and slope geometries,
horizontal ground accelerations predicted for the 2005 Kashmir earthquake by the four GMPEs,
and k
h
= 0.5, with the LC data from Sato et al. (2007). The static FS was calculated to be
approximately 2.6. When the horizontal acceleration was incorporated, the FS becomes 1.3 at
distances close to the fault. The mean FS did not become smaller than unity even when the

99

horizontal acceleration was at its maximum. Although some combinations of material properties
and slope geometries cause the FS to become smaller than unity, this mean FS distribution
cannot explain the high LC near the fault. Similar disparity between the LC data and pseudo-
static FS was observed when using mean strength and geometry parameters for the 1989 Loma-
Prieta, U.S.; 1999 Chi-Chi, Taiwan; and 2008 Wenchuan, China, earthquakes.
6.6 Summary
This chapter summarizes the general approaches to analyze seismic slope stability.
Terzaghi (1950) first applied a pseudo-static approach to analyze seismic slope stability that
results in prediction of factor of safety against sliding. The Newmark (1965) sliding block
analysis is to estimate permanent displacements of a sliding mass. Makdisi and Seed (1978)
proposed a simple method for the seismic design of small embankments that uses the concept
originally proposed by Newmark (1965). Many researchers later modified the Newmark method
to deal with compliant slope materials (e.g., Lin and Whitman 1983; Kramer and Smith 1997;
Rathje and Bray 2000). In order to evaluate slope failure including flow failures from their
initiation to termination, Kokusho and Kabasawa (2003) proposed an energy approach. The
pseudo-static slope stability analysis is selected in this study to evaluate the regional
performance of slopes during earthquake.
A pseudo-static slope stability analysis was conducted for the 2005 Kahsmir, Pakistan,
earthquake. The rock strength parameters and slope geometry parameters were obtained from the
literature (Sato et al. 2007; Kamp et al. 2008; Owen et al. 2008). Using horizontal ground
accelerations predicted using four GMPEs released by Next Generation Attenuation (NGA)
projects (Abrahamson and Silva 2008; Boore and Atkinson 2008; Campbell and Bozorgnia 2008;
Idriss 2008), mean factor of safety against sliding was computed. However, this FS could not
capture the trend of LC data. The static FS was calculated to be approximately 2.6. When the
horizontal acceleration was incorporated, the FS becomes 1.3 at distances close to the fault, but
is still greater than unity.


100

Table 6.1 Characteristics of earthquake-induced landslides (disrupted type) (Keefer 2002).
Name Type of movement
Internal
disruption
Water content
D U PS S
Typical
depths
Min.
slope ()
Typical
velocities
Typical volumes Typical displacements
Rock falls
Bouncing, rolling,
free fall
High or
very high
Shallow 40
Extremely
rapid
Most less than 1 10
4
m
3
;
maximum reported 2
10
7
m
3

May fall to base of steep
source slope and move as far
as several tens or hundreds of
meters farther, on relatively
gentle slopes
Disrupted
rock slide
Translational sliding High Shallow 35
Rapid to
very rapid
Most less than 1 10
4
m
3
;
maximum reported 2
10
9
m
3

May slide to base of steep
source slope and several tens
or hundreds of meters farther,
on relatively gentle slopes
Rock
avalanches
Complex, involving
sliding,
flow, and
occasionally free fall
Very high Deep 25
Very rapid
to
extremely
rapid
5 10
5
2 10
8
m
3
or
more
Several kilometers
Soil falls
Bouncing, rolling,
free fall
High or
very high
Shallow 40
Extremely
rapid
Most less than 1,000 m
3
;
maximum volumes not
well documented
Most come to rest at or near
bases of steep source slopes
Disrupted
soil slides
Translational sliding High Shallow 15
Moderate
to rapid
Most less than 1 10
4
m
3
;
maximum reported 4.8
10
7
m
3

May slide to base of steep
source slope and several tens
or hundreds of meters farther,
on relatively gentle slopes
Soil
avalanches
Complex, involving
sliding,
flow, and
occasionally free fall
Very high Shallow 25
Very rapid
to
extremely
rapid
Volumes not well
documented; maximum
reported 1.5 10
8
m
3

Several tens of meters to
several kilometers beyond
steep source slopes
Notes: Names: rock signifies bedrock that is relatively firm and intact prior to landslide initiation, and soil signifies loose, unconsolidated or poorly
cemented aggregates of particles that may or may not contain organic materials. Internal disruption: slight signifies landslide consists of one or a few coherent
blocks; moderate signifies several coherent blocks; high signifies numerous small blocks and individual soil grains and rock fragments; very high signifies
nearly complete disaggregation into individual soil grains or small rock fragments. Depth: shallow signifies generally < 3 m deep; deep signifies generally > 3 m
deep. Water content: D = dry; U = moist but unsaturated; PS = partly saturated; S = saturated. Velocity: very slow = 1 106
3
10
6
m/min; slow = 3 10
6
3
10
5
m/min; moderate = 3 10
5
0.001 m/min; rapid = 0.0010.3 m/min; very rapid = 0.3180 m/min; extremely rapid = >180 m/min.



101

Table 6.2 Characteristics of earthquake-induced landslides (coherent type) (Keefer 2002).
Name
Type of
movement
Internal
disruption
Water
content
D U PS S
Typical
depths
Min.
slope
()
Typical
velocities
Typical volumes Typical displacements
Rock
slumps
Rotational
sliding
Slight or
moderate
? Deep 15 Slow to rapid
Most between 100 and a
few million m
3
; maximum
at least tens of millions of
m
3

Typically less than 10
m; occasionally 100 m
or more
Rock block
slides
Translational
sliding
Slight or
moderate
? Deep 15 Slow to rapid
Most between 100 and a
few million m
3
; maximum
at least tens of millions of
m
3

Typically less than 100
m; maximum
displacements not well
documented
Soil slumps
Rotational
sliding
Slight or
moderate
? Deep 7 Slow to rapid
Most between 100 and 1
10
5
m
3
; occasionally 1
10
5
to several million m
3

Typically less than 10
m; occasionally 100 m
or more
Soil block
slides
Translational
sliding
Slight or
moderate
? ? Deep 5 Slow to rapid
Most between 100 and 1
10
5
m
3
; maximum reported
1.12 10
8
m
3

Typically less than 100
m; maximum
displacements not well
documented
Slow earth
flows
Translational
sliding and
internal flow
Slight
Generally
shallow;
occasionally
deep
10
Very slow to
moderate;
occasionally,
with very rapid
surges
Most between 100 and 1
10
6
m
3
; maximum reported
between 3 10
7
and 6
10
7
m
3

Typically less than 100
m; maximum
displacements not well
documented
Lateral
spreads and
flows
Soil lateral
spreads
Translation on
fluid basal zone
Generally
moderate;
occasionally
slight or high
Variable 0.3 Very rapid
Most between 100 and 1
10
5
m
3
; largest reported 9.6
10
6
m
3

Typically less than
10m; maximum
reported 600 m
Rapid soil
flows
Flow Very high ? ? ? Variable 2.3
Very rapid to
extremely rapid
Volumes not well
documented; largest are at
least several million m3
A few m to several km
Subaqueous
landslides
Generally lateral
spreading or
flow;
occasionally
sliding
Generally high
or very high;
occasionally
moderate or
slight
Variable 0.5
Generally rapid
to extremely
rapid;
occasionally
slow to moderate
Volumes not well
documented; largest are at
least tens of millions of m
3

Not well documented,
but some move more
than 1 km
Note: See Table 6.1.


102

Table 6.3 Minimum M
L
required to trigger landslides (Keefer 1984). The numbers in parentheses
are M
w
converted using conversion from Grunthal and Wahlstrom (2003).
M
L
(M
w
) Description
4.0 (3.6) Rock falls, rock slides, soil falls, disrupted soil slides
4.5 (4.1) Soil slumps, soil block slides
5.0 (4.6)
Rock slumps, rock block slides, slow earth flows, soil
lateral spreads, rapid soil flows, and subaqueous landslides
6.0 (5.7) Rock avalanches
6.5 (6.3) Soil avalanches



103

Table 6.4 Hoek-Brown parameters and geometry parameters for four cases.
Parameter
Selected values
Notes References
2005
Kashmir
1989
Loma-
Prieta
1999
Chi-Chi
2008
Wench-
uan

(kN/m
3
)
21 1 21 1 24 1 24 1
Estimated from rock
type and degree of
weathering
Hoek and Bray
(1981)
o
ci
(MPa) 2 0.8 2 2
Estimated from rock
type and degree of
weathering
Hoek et al. (1998)
m
i
14 14 14 14
Estimated from rock
type
Hoek et al. (1998)
GSI 30 30 30 30
Estimated from rock
type, rock mass
structure, and joint
surface conditions
Marinos et al.
(2005)
DF 0 0 0 0
No evidence of
blasting or stress relief
Hoek et al. (1998)
|'
()
40 6 33 5 51 8 40 6
Estimated using Hoek-
Brown criterion
Hoek et al. (2002)
c'
(kPa)
27 8 20 6 11 3 27 8
Estimated using Hoek-
Brown criterion
Hoek et al. (2002)
o
t
(kPa)
0.73
0.15
0.29
0.07
0.73
0.15
0.73
0.15
Estimated using Hoek-
Brown criterion
Hoek et al. (2002)
D
t

(m)
3 0.67 3 0.67 0.7 0.2 3 0.67
Based on generalized
description of typical
depths of disrupted
slides in rock slopes
Keefer (1984)
Khazai and Sitar
(2004)
Owen et al. (2008)
Yin et al. (2010)
i
()
28 9 27 9 65 15 33 10 See Figure 6.5
Khazai and Sitar
(2004)
Kamp et al. (2008)
Yin et al. (2010)






104


Figure 6.1 Relations between area affected by landslides and earthquake magnitude (Keefer
2002). Circles are data from Keefer (1984) and Keefer and Wilson (1989). Dashed line is
approximate upper bound from Keefer (1984). Solid line is least-squares linear regression mean
from Keefer and Wilson (1989). Most magnitudes smaller than 7.5 are surface-wave (M
s
), and
most magnitude of 7.5 or larger are moment magnitude (M
w
).



105


Figure 6.2 Maximum distance from fault rupture zone to landslides with respect to earthquake
magnitudes. Dashed line represents disrupted slides and falls, dash-double-dot line represents
coherent slides, and dotted line represents lateral spreads and earth flows (Keefer 1984).

106


Figure 6.3 Distribution of landslides triggered by the 2005 Kashmir earthquake, shown as red
dots. Epicenter (star mark), Balokot-Garhi fault (light line), Jhelum fault (dark line), and cities
of Muzaffarabad (M) and Balakot (B) are also shown. (Sato et al. 2007).

107


Figure 6.4 Photograph of shallow landslides triggered by 2005 Kahsmir, Pakistan, earthquake
(by Prof. Hashash).


108


Figure 6.5 Distribution of slope angles for (a) 2005 Kashmir, Pakistan, earthquake (Kamp et al.
2008), (b) 1989 Loma Prieta earthquake (Khazai and Sitar 2004), (c) 1999 Chi-Chi, Taiwan,
earthquake (Khazai and Sitar 2004), and (d) 2008 Wenchuan, China, earthquake (Yin et al.
2010).


109


Figure 6.6 Horizontal accelerations with respect to source-to-site distance predicted using the
NGA GMPEs for the M
w
7.6 Kashmir, Pakistan, earthquake.

0 10 20 30 40 50
Distance from surface projection of updip edge of fault plane (km)
0
0.2
0.4
0.6
0.8
1
1.2
H
o
r
i
z
o
n
t
a
l

g
r
o
u
n
d

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Abrahamson-Silva (2008)
Boore-Atkinson (2008)
Campbell-Bozorgnia (2008)
Idriss (2008)

110


Figure 6.7 Landslide concentration data for the 2005 Kashmir, Pakistan, earthquake by Sato et al.
(2007) and the mean factor of safety calculated by using existing pseudo-static method.

0 10 20 30 40 50
Distance from surface projection of updip edge of fault plane (km)
0
0.4
0.8
1.2
1.6
2
L
a
n
d
s
l
i
d
e

c
o
n
c
e
n
t
r
a
t
i
o
n
,

L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
2.4
2.6
2.8
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
LC for 2005 Kashmir, Pakistan earthquake
FS with k
h
Static FS
Mean parameters
Tensile strength = 0.73 kPa
Unit weigth = 21 kN/m
3
Cohesion intercept = 27 kPa
Friction angle = 40
o
Slope thickness = 3 m
Slope angle = 28
o
22

111

CHAPTER 7. FACTORS THAT AFFECT PSEUDO-STATIC SLOPE STABILITY
7.1 Introduction
To address the disparity between LC data and the FS predicted by existing pseudo-static
slope stability approach, this study proposes three factors that potentially would affect LC near
earthquake source zones in addition to the horizontal acceleration: vertical ground acceleration,
topographic effects, and bond break effect. This chapter introduces the vertical GMPE
proposed by Campbell and Bozorgnia (2003) and Ambraseys et al. (2005b) and vertical-to-
horizontal ratios proposed Ambraseys and Simpson (1996) and Elnashai and Papazouglou (1997).
The field and numerical studies evaluating topographic effects are also presented. Finally, the
bond break effect associated with vertical acceleration is proposed.
7.2 Incorporating vertical acceleration into pseudo-static analysis
Vertical accelerations traditionally have been ignored in pseudo-static slope stability
analyses because it is believed to have little influence on slope stability because it reduces both
the driving and resisting forces. For a typical set of strength and geometric properties, Figure 7.1
shows that the vertical pseudo-static coefficient (k
v
) actually increases the FS when k
h
is smaller
than 0.2. However, when k
h
exceeds 0.2, k
v
decreases the FS. In addition, k
v
tends to increase the
FS when the sliding mass is thinner and the slope is steeper, regardless of k
h
, while k
v
decreases
the FS when the sliding mass is thicker and the slope is gentler (assuming constant strength
parameters with depth).
In addition to the direct effect of k
v
, many researchers have reported that the maximum
vertical acceleration can equal or exceed the horizontal acceleration close to the causative fault,
particularly for large magnitude earthquakes. To evaluate vertical accelerations, some
researchers, e.g, Campbell and Bozorgnia (2003) and Ambraseys et al. (2005b), have proposed
vertical acceleration attenuation relationships; however, these methods are limited by a lack of
recorded ground motions. Despite this limitation, they used available strong-motion data for
shallow crustal earthquakes recorded throughout the world (Campbell and Bozorgnia 2003) and
in Europe and the Middle East (Ambraseys et al. 2005b). Other researchers have proposed
relationships between horizontal and vertical accelerations. For example, Figure 7.2 presents
vertical-to-horizontal acceleration ratios (V/H) proposed by Ambraseys and Simpson (1996) and
Elnashai and Papazouglou (1997). The Ambraseys and Simpson (1996) relation is based on 110

112

near-field recordings (R < 15 km) relatively large (surface wave magnitude, Ms > 6), shallow
(focal depth < 20 km), interplate earthquakes. They suggested that V/H slightly exceeds 1 for
source-to-site distances less than about 5 km for large thrust fault ruptures and for moderate-to-
strong strike-slip events. Their relations for reverse fault were linearly extrapolated to a distance
of 50 km. Elnashai and Papazouglou (1997) proposed relations that adopted Ambraseys and
Simpson (1995) for distances to the fault less than 15 km, Abrahamson and Litehiser (1989) for
distances larger than 30 km, and a linear interpolation for distances between 15 km and 30 km.
(see Figure 7.2). These V/H relations represent free-field conditions, as most records were
obtained from free-field stations, with some from basements or ground floors of buildings. Both
relations suggest that the V/H ratio slightly exceeded 1.0 for M
s
7.5 earthquakes in the near-
field. Due to a lack of studies on the vertical pseudo-static coefficient k
v
, this study used the same
scaling factor as the horizontal coefficient. i.e., k
v
= 0.33, 0.5, 0.9(a
v,max
/g).
7.3 Topographic effects
It is well-known that seismic waves can be amplified or deamplified when propagated
through a soil column, depending on soil properties. These effects are termed one-dimensional
(1D) site effects. Seismic waves also can be amplified or deamplified due to constructive or
destructive interference when the waves encounter surface topographic irregularities such as
valleys, peaks, and plateaus. These phenomena are termed topographic effects or two-
dimensional (2D) site effects.
In many settings, waves constructively interfere near the crest of a slope. For example,
during the 1971 San Fernando, California, earthquake (M
w
6.6), severe ground cracking was
observed in a very restricted area at the top of a ridge while no significant slope failures occurred
(Nason 1971) (Figure 7.3 a). Similarly, Castellani et al. (1982) observed that during the Irpinia,
Italy, earthquake (M
w
6.9), building damage was concentrated around the top of a hill, while the
village at the foot of the hill was only slightly damaged (Figure 7.3 b).
These observations led to several full-scale field studies evaluating topographic effects
where aftershocks were monitored along hills or ridges (e.g., Celebi 1987; Hartzell et al. 1994;
Graizer 2009). Recently Hough et al. (2010) investigated topographic effects using aftershocks
of the 2010 M
w
7.0 Haiti earthquake. However, in many of these cases 1D site effects cannot be
conclusively distinguished from topographic effects. Therefore, some laboratory shaking table

113

and centrifuge studies have been performed to evaluate topographic effects (e.g., Kovacs et al.
1971; Stamatopoulos et al. 2007).
Many numerical studies have been conducted to investigate topographic effects,
including studies by Idriss and Seed (1967), Boore (1972), Smith (1975), Sanchezsesma et al.
(1982), Sitar and Clough (1983), Ashford et al. (1997), Ashford and Sitar (1997), and
Bouckovalas and Papadimitrious (2005). Many recent studies have addressed specific
observations or recordings. For example, Gazetas et al. (2002), Assimaki and Gazetas (2004),
and Assimaki et al. (2005) evaluated topographic effects observed during the 1999 Athens,
Greece earthquake (M
w
5.9). Lungarini et al. (2005) used a finite element method to simulate
topographic effects observed at Mt. Etna, Italy and observed that the maximum computed
displacement occurred at or near the peak of the mountain.
The topographic horizontal amplification factors observed in these field and numerical
studies are summarized in Figure 7.4. Results reported in terms of velocity and displacement
were not included in Figure 7.4. Some studies defined the amplification factor as the ratio of
a
h,max
at crest to a
h,max
at free field beyond the crest. Others used the ratio of a
h,max
at the crest to
a
h,max
at the toe. No distinction was made between these two definitions of amplification factors
assuming that a
h,max
at free filed beyond the crest and a
h,max
at toe are the same for a rigid
homogeneous rock. Figure 7.4 also includes the topographic amplification factors recommended
by the European Seismic Code provision (EC8 2000) and French Seismic Code provision
(French Association for Earthquake Engineering (AFPS) 1995).
Some of the studies mentioned above also examined topographic amplification factors for
vertical accelerations (e.g., Idriss and Seed 1967; Smith 1975; Sitar and Clough 1983; Ashford
and Sitar 1997; Ashford et al. 1997; Assimaki et al. 2005; Bouckovalas and Papadimitriou 2005).
However, all of these studies considered the response of vertical acceleration that was
transformed from a horizontal input motion (i.e., no vertical input acceleration was included).
Celebi (1991) reported vertical topographic amplification factors from 0.91 to 1.86 during
aftershocks of the 1983 Coalinga, California, earthquake (M
w
=6.5). Based on numerical
simulations, Zhao and Valliappan (1993) reported topographic amplification factors of 1.5 to 1.8
for the response of various topographies to vertically-incident P-waves, These values are
reasonably consistent with the horizontal topographic amplification factors reported in Figure 7.4;

114

therefore, vertical topographic amplification factors were assumed to be same as horizontal
topographic amplification factors.
Most of the shallow, disrupted landslides examined in this study were initiated near slope
crests, topographic amplification factors only near the crest were considered. As illustrated in
Figure 7.4, horizontal topographic amplification factors range from about 1.0 to 1.9 at the crest.
To account for variability, the three-sigma rule was applied, and a mean of 1.45 and a standard
deviation of 0.15 were estimated. Assuming that vertical topographic amplification is same as
the horizontal amplification, horizontal and vertical topographic factors of 1.3, 1.45, and 1.6
were used in the statistical analyses (described subsequently) with weights of 0.2, 0.6, and 0.2,
respectively. The range also is supported by Ashford and Sitar (2002) who suggested that
amplification due to topography is on the order of 50% regardless of slope geometries and soil
profiles.

7.4 Potential bond break effect associated with vertical acceleration
Figure 7.5 (a) schematically illustrates the forces involved in a static infinite slope (side
forces were assumed to be equal and opposite, and therefore were not included). When an
earthquake occurs, the P-waves result in vertical surface accelerations and potentially can break
weak bonds (i.e., cementation or cohesion) if a pseudo-static vertical acceleration exceeds the
slice weight plus tensile force along the base of a slice (Figure 7.5 b). When a distance from the
fault is less than 5 km, horizontal and vertical peak ground accelerations can be coincident
(Collier and Elnashai 2001). In this case, it was considered that the vertical accelerations still
contribute to bond breakage at the same time the horizontal and vertical accelerations shake the
sliding mass. Furthermore, high amplitude accelerations oriented perpendicular to the sliding
mass, due to the vector resultant of horizontal and vertical accelerations, can increase the
possibility of bond break. However, this effect was not considered in the analysis.
Using force equilibrium (Figure 7.5 b), bond break occurs when:

W i l F
t v
+ > cos o Eq. (7-1 a)


115

1
cos
max ,
+

>
t
t
v
D
i
g a

o

Eq. (7-1 b)

where F
v
is the vertical pseudo-static force, equal to k
bond
a
v,max
W. The term k
bond
, is taken as
1.0 based on the assumption that the maximum acceleration will progressively occur at a slope
and potentially break any bond forces along the entire base of the sliding mass. As discussed
earlier, topographic amplification factors only near the crest were considered because most of the
shallow, disrupted landslides examined in this study were initiated near slope crests. If the bond
forces within the rock mass are broken, rock cohesion is considered to be zero for pseudo-static
stability calculations.
7.5 Summary
Vertical accelerations traditionally have been ignored in pseudo-static slope stability
analyses because it is believed to have little influence on slope stability because it reduces both
the driving and resisting forces. However, this study revealed that the vertical accelerations play
a role in decreasing the FS with large horizontal accelerations. In addition, Ambraseys and
Simpson (1996) and Elnashai and Papazouglou (1997) proposed vertical-to-horizontal
acceleration ratios that predict the vertical acceleration equal to or greater than horizontal
acceleration close to the fault, particularly for large magnitude earthquakes. Furthermore, the
vertical acceleration contributes to break the bond force within the rock mass, especially when it
is amplified by topographic effect. Based on numerous field and numerical studies, the
topographic amplification factors of 1.3, 1.45, and 1.65 were proposed with weights of 0.2, 0.6,
and 0.2, respectively.



116


Figure 7.1 Pseudo-static FS computed using various combinations of horizontal and vertical
pseudo-static coefficient, k
h
and k
v
. Strength and geometric properties represent typical values
listed in Table 6.4.


0 0.1 0.2 0.3 0.4 0.5
Horizontal pseudo-static coefficient, k
h
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
2.4
2.6
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
k
v
= 0.5
k
v
= 0.4
k
v
= 0.3
k
v
= 0.2
k
v
= 0.1
k
v
= 0
Unit weight = 23 kN/m
3
Friction angle = 40
o
Cohesion intercept= 22 kPa
Slope angle = 30
o
Slope thickness = 3 m

117


Figure 7.2 Vertical-to-horizontal acceleration ratio for thrust faults by Elnashai and Papazouglou
(1997) and Ambraseys and Simpson (1996). Extrapolated portion from Ambraseys and Simpson
(1996) is shown in thinner lines.


0 10 20 30 40 50
Distance from the fault (km)
0.2
0.4
0.6
0.8
1
1.2

V
e
r
t
i
c
a
l

a
c
c
e
l
e
r
a
t
i
o
n
H
o
r
i
z
o
n
t
a
l

a
c
c
e
l
e
r
a
t
i
o
n
Elnashai and Papazouglou (1997)
Ambraseys and Simpson (1996)
Ms=7.5
Ms=7.0
Ms=6.5
Ms=6.0
Ms=5.5
M
s
=
7
.
5
M
s
=
7
.
0
M
s
=
6
.
5
M
s
=
6
.
0
M
s
=
5
.
5

118


Figure 7.3 (a) Ground cracking (indicated by XXXs) during the 1971 San Fernando, California,
earthquake (Nason 1971); (b) Destruction of a village during the Irpinia, Italy, earthquake
(Castellani et al. 1982).
(a)
(b)

119



Figure 7.4 Topographic factors proposed by various researchers.


120


Figure 7.5 Illustration showing procedure of slope stability analysis taking account bond break
effect. (a) Dimensions and forces involved in static stability, where W = slice weight; N = normal
force; and T = shear force. Side forces are assumed to be equal and opposite, and therefore were
not included. (b) Evaluation of bond break where a
v,max
= maximum vertical acceleration; o
t
=
tensile strength.


l
D
W
T
N
W
o
t
lcosi
i
F
v
=a
v,max
/gW
(a) (b)

121

CHAPTER 8. PROPOSED NEW APPROACH TO PSEUDO-STATIC SLOPE
STABILITY ANALYSIS AND VERIFICATION USING EARTHQUAKE CASES
8.1 Introduction
The proposed procedure for pseudo-static slope stability analysis is presented in Figure 8.1.
This procedure accounts for potential vertical accelerations, topographic effects, and bond break
effects which were discussed in Chapter 7. This chapter compares factors of safety predicted
using the proposed pseudo-static slope stability analysis to landslide concentration data from four
earthquakes (2005 Kahsmir, Pakistan, earthquake; 1989 Loma-Prieta, U.S., earthquake; 1999
Chi-Chi, Taiwan, earthquake; and 2008 Wenchuan, China, earthquake). A logic tree approach
was employed to perform all the calculations incorporating the uncertainties in each of the
parameters.
8.2 2005 Kashmir, Pakistan earthquake
As mentioned earlier, vertical accelerations were predicted using two attenuation
relationships (Campbell and Bozorgnia 2003; Ambraseys et al. 2005b) and two vertical-to-
horizontal acceleration ratios (Ambraseys and Simpson 1996; Elnashai and Papazoglou 1997).
These predicted vertical accelerations for the 2005 Kashmir earthquake are shown in Figure 8.2.
Table 6.4 summarizes the geometries and rock properties used in the analysis. Figure 8.3 shows
the logic tree employed for the 2005 Kashmir earthquake case. The logic tree incorporates mean
() values and mean values 1 standard deviation () for material properties (, c', |', and o
t
),
and slope geometry parameters (i and D
t
). For subsequent statistical analysis, it was assumed that
these parameters followed normal distributions, and the weights of 0.6 and 0.2 were assigned to
and values, respectively as shown in Figure 8.3. The logic tree incorporates the four
horizontal and four vertical acceleration attenuation relations described above, three topographic
factor values, and three values of horizontal and vertical pseudo-static coefficients. This results
in 314,928 branches in the logic tree. The logic tree allows us to calculate FS for each branch
individually and estimate the mean FS using the weights assigned to each branch.
Figure 8.4 compares the mean FS to the LC data collected by Sato et al. (2007). The
mean static FS was calculated to be about 2.6, and the mean FS decreased to 1.3 near the fault
when horizontal acceleration was included. Clearly these FS values cannot explain the observed
LC data. Horizontal accelerations with topographic effects also did not decrease the FS below

122

unity. However, when topographic amplification factors were applied to both horizontal and
vertical accelerations, the FS became less than unity within about 10 km from the fault. When
bond break effects were applied, the FS became smaller than unity at R ~ 13 km. Both of these
effects are consistent with the increase in LC that occurs around R ~ 15 km.
Figure 8.4 also includes the FS 1 standard deviation. Given the large number of
variables in the logic tree, the considerable uncertainty in FS was unavoidable. To illustrate the
effects of individual variable uncertainty, Figure 8.5 presents mean FS and mean FS 1o for
several variables involved in the FS calculation. The standard deviation for the static FS is about
0.89. However, when vertical acceleration, topographic effect, and bond break effect were
included in the analysis, the standard deviation decreased to 0.63. Thus, the uncertainties in the
pseudo-static FS calculation do not appear to overshadow the observed trends in mean FS.
Similarly, Figure 8.6 shows the results of a sensitivity analysis for the following rock and
geometric parameters: , |', c', i, and D
t
. Since the failed rock slopes in this study are consistently
highly weathered, the tensile strengths are estimated to be very small. Therefore, the uncertainty
in tensile strength did not have much effect on FS in this study, thus the result for tensile strength
is not shown here. It is notable that the variation in slope angle results in the largest standard
deviation, whereas the variation of unit weight produces minor standard deviation. Another
observation is that variation in FS resulting from rock and geometric parameter uncertainties
decreases as R decreases. This occurs because ground accelerations have more influence on the
FS as ground accelerations become larger. Figure 8.7 shows the results of a sensitivity analysis
considering horizontal and vertical pseudo-static coefficients, horizontal and vertical ground
accelerations, and the topographic factor.
Monte-Carlo simulations were conducted to validate the logic tree approach. Truncated
normal distributions (truncated at 3o) were selected for the rock parameters and slope thickness.
For slope angles, the same distributions that are reported as shown in Figure 6.5 were used for all
earthquake cases. The simulations involved 10,000 runs. Figure 8.8(a) shows the mean FS and
mean FS 1 estimated using the Monte-Carlo simulations for the 2005 Kashmir earthquake.
The standard deviation for FS was estimated to be slightly larger than that generated by the logic
tree approach. This occurs because the Monte-Carlo simulations considered wider ranges for
each parameters (up to three standard deviations of each mean parameter), whereas the logic tree
approach constrains the parameters to within the mean one standard deviation. However, the

123

overall results are similar to those by the logic tree approach. When all factors (horizontal and
vertical accelerations, topographic effects, and bond break effects) were considered, the factor of
safety became less than unity at approximately 14 km from the updip edge of the fault plane,
close to the distance where the LC data start to increase abruptly.
8.3 1989 Loma-Prieta, U.S. earthquake
In 1989, a M
w
6.9 earthquake occurred in the San Francisco Bay-Monterey Bay region of
California. The rupture occurred on the San Andreas Fault system with a rupture depth of 3 km
to 18 km. The faulting mechanism was oblique reverse with a dip of 60 SW. The region
affected by the earthquake includes a variety of sedimentary, igneous, and metamorphic rocks
and unconsolidated sedimentary deposits from Jurassic through Holocene-aged (Keefer and
Manson 1998). Keefer (2000) performed a statistical analysis of the distribution of 1046
landslides in the southern Santa Cruz Mountains, triggered by the 1989 Loma-Prieta earthquake
over an area of 192.41 km
2
. The fault and landslide locations are shown in Figure 8.11. Based on
these data, Keefer (2000) expressed the landslide distribution as a landslide concentration (LC),
which is defined as the number of landslides per square kilometer of surface area. Approximately
74% of the 1046 landslides studied by Keefer (2000) were disrupted slides and falls, and the
majority of the 1046 landslides occurred in the Purisima Formation, a weakly-cemented
sandstone and siltstone.
Keefer (2000) reported that the Purisima Formation in which most of landslides occurred
consists of weakly cemented sandstones and siltstones, and the rocks are poorly to moderately
indurated, and structurally deformed by pervasive folding and local faulting. Keefer and Manson
(1989) also reported that disrupted landslides in all units occurred in materials that typically were
weakly cemented, closely fractured, intensely weathered, and broken by joints. Based on these
descriptions, Table 6.4 summarizes the Hoek-Brown parameters assigned for this case. Mohr-
Coulomb strength parameters were computed as: |' = 33 and c' = 20 kPa for the appropriate
range of stresses (corresponding to D
t
). These calculated values are consistent with the mean |' =
33.1 and median c' = 25.2 kPa measured by McCrink and Real (1996) in the Purisima
Formation at a depth of 3 m.
As discussed by Keefer (1984), disrupted slides and falls typically involve only the upper
few meters of overburden. Therefore, D
t
= 1 to 5 m were considered. Figure 6.5 (b) shows the

124

distribution of slope angles for slope that failed during the 1989 Loma Prieta earthquake (Khazai
and Sitar 2004). A normal distribution was assumed with a mean slope angle of 27 and a
standard deviation of 9 to represent the failed slope data. Table 6.4 summarizes the geometric
parameters for this case.
Figure 8.11 shows the results of the pseudo-static infinite slope stability analyses using
the proposed procedure and logic tree approach. The horizontal acceleration (which increases for
shorter source-to-site distances, R) clearly reduces the FS, but is not sufficient to drop the mean
FS below unity. As anticipated, the vertical acceleration alone does not have much effect on FS
because M
w
< 7.0. Incorporating topographic effects on the horizontal acceleration significantly
reduces the FS; however, the mean FS does not drop below unity R 6 km. This is inconsistent
with the LC data, which shows a significant change in landslide concentration at R ~ 9 km.
Incorporating topographic effects with the vertical acceleration also decreases the FS at near-
fault distances, with the mean FS = 1.0 at R ~ 10 km, consistent with the significant change in
LC. When the bond break effect is incorporated, a further slight decrease in FS occurs at R 3
km. Similar to the case of Kashmir earthquake, Monte-Carlo simulations yielded results similar
to the logic tree approach, as shown in Figure 8.8(b).
Figure 8.12 presents mean FS and mean FS 1o for several variables involved in the FS
calculation. Similarly with the case for the 2005 Kahsmir, Pakistan, earthquake, the standard
deviation decreases as more factors are included. The standard deviation for the static FS is about
0.73. However, when all other factors were considered, the standard deviation decreased to 0.44
near fault. Figure 8.13 and Figure 8.14 show the results of a sensitivity analysis for the rock and
geometric parameters. The sensitivity analyses again show the similar trend with the case or the
2005 Kahsmir, Pakistan, earthquake.
8.4 1999 Chi-Chi, Taiwan earthquake
The 1999 Chi-Chi earthquake (M
w
7.6) occurred in the mountainous terrain of central
Taiwan, and triggered more than 10,000 landslides throughout an area of approximately 11,000
km
2
(Hung 2000). The Cher-Long-Pu Fault that ruptured during this event is a thrust fault (dip =
38, depth to the bottom of rupture = 20 km) trending from south to north (Meunier et al. 2007)
as shown in Figure 8.15. The horizontal and vertical accelerations predicted for this earthquake
are shown in Figure 8.16.

125

Figure 8.15 also shows the landslide locations along the fault and the geologic setting of
the affected area. Most of the landslides were shallow, disrupted slides and falls in Tertiary
sedimentary and submetamorphic rocks (Khazai and Sitar 2004). The landslide masses
commonly were dry, highly disaggregated, weakly cemented, closely fractured, intensely
weathered, and broken by conspicuous, highly persistent joints (Khazai and Sitar 2004).
Estimated rock properties are presented in Table 6.4. Khazai and Sitar (2004) reported that depth
of sliding typically ranged from several decimeters to a meter, and Figure 6.5(c) presents the
range of slope angles where the landslides occurred. Although there is slight difference between
the distribution of documented slopes and the theoretical normal distribution, it is considered that
this difference is acceptable. The resulting geometric parameters are also summarized in Table
6.4. Using the geometric and rock properties in Table 6.4, the computed Mohr-Coulomb strength
paremeters are: |' = 51 and c' = 11 kPa for the appropriate range of stresses (corresponding to
D). Unlike other cases, the 1999 Chi-Chi, Taiwan earthquake triggered the landslides mostly on
steep slopes. In shallow slides, wave incoherency is more likely on steep slopes than on gentle
slopes because the vertical travel path of the ground motions increases as the slopes become
steeper. It was also found that the FS becomes unreasonable when both k
v
and i are large in Eq.
6-2. Therefore, to account for potential wave incoherency, yet avoid unreasonable estimates of
FS, k
h
and k
v
were limited to values of 0.7(a
h,max
/g) and 0.7(a
v,max
/g), respectively, for i 50.
Figure 8.17 shows the mean FS compared to LC data collected by Khazai and Sitar
(2004). As a result of the fairly steep slopes, the static FS was computed as 1.4. The FS with the
horizontal acceleration became smaller than unity at R ~ 27 km, which is consistent with an
increase in LC. However, the FS increases when the vertical accelerations are included because
the sliding mass is thin and the slope is steep. When topographic effects were considered, the FS
decreases to unity for R ~ 30 km, which approximately matches the distance where LC begins to
increase. Bond break effects occur at R ~ 13 km, and the mean FS decreases significantly near
the fault. However, it is anticipated that the LC data does not show a significant increase near the
fault (even with the small FS at R < 10 km) because the slopes become gentler and more sparsely
located near the fault than at some distance from the fault (Meunier et al. 2007). Similar to the
case of Kashmir earthquake, Monte-Carlo simulations yielded results similar to the logic tree
approach, as shown in Figure 8.8(c).

126

Figure 8.18 presents mean FS and mean FS 1o for several variables involved in the FS
calculation. Figure 8.19 and Figure 8.20 show the results of a sensitivity analysis for the rock and
geometric parameters.
8.5 2008 Wenchuan, China earthquake
On 12 May 2008, the Wenchuan earthquake (M
w
7.9) occurred in the Sichuan Province
of China. The earthquake occurred in the compression zone between the Indian and Eurasian
plates along the Longmenshan-Beichuan-Yinxiu fault zone (rupture length = 300 km, rupture
depth = 14 to 19 km, dip = 60) (Cui et al. 2009). The isoseismic map along the fault is shown in
Figure 8.21. The horizontal and vertical accelerations predicted for this earthquake are shown in
Figure 8.22.
Figure 8.21 shows the distribution of over 56,000 earthquake-triggered landslides,
distributed over an area of 811 km
2
. These were identified by Dai et al. (2011) using aerial
photos and high-resolution satellite images. Yin et al. (2009) and Dai et al. (2011) reported that
most of the landslides occurred in weathered or heavily fractured sandstone, siltstone, slate, and
phyllite. Based on these descriptions, Table 6.4 summarizes the estimated rock properties. The
majority of landslides were shallow, disrupted landslides and rock falls from steep slopes,
typically involving the top few meters of weathered bedrock, regolith, and colluvium (Dai et al.
2011). Similarly, Yin et al. (2009) reported typical sliding depths of about 1 to 5 m and slope
angles as shown in Figure 6.5. From these descriptions, Table 6.4 includes the estimated
geometric parameters for these landslides. The equivalent Mohr-Coulomb strength parameters
were calculated as |' = 40 and c' = 27 kPa for the appropriate range of stresses.
Figure 8.23 compares the mean FS to the LC data collected by Dai et al. (2011). The
static FS was calculated to be 2.1, and the FS decreased to about 1.1 near the fault when
horizontal accelerations were included. This mean FS is not consistent with the LC data,
particularly at short source-to-site distances. When topographic effects were applied to horizontal
accelerations, the mean FS becomes smaller than unity at R ~ 10 km. However, a slight decrease
in FS within 10 km from the fault cannot explain the large LC in the near field. Topographic
effects on both horizontal and vertical accelerations, as well as bond break effects, decrease the
FS below unity within about 13 km of the fault, consistent with the abrupt change in LC at R ~

127

12 km. The Monte-Carlo simulations yielded similar results to the logic tree approach, as shown
in Figure 8.8(d).
Figure 8.24 presents mean FS and mean FS 1o for several variables involved in the FS
calculation. Figure 8.25 and Figure 8.26 show the results of a sensitivity analysis for the rock and
geometric parameters.
8.6 Summary and conclusion
This chapter predicts FS using the proposed pseudo-static slope stability analysis for four
earthquakes (2005 Kahsmir, Pakistan, earthquake; 1989 Loma-Prieta, U.S., earthquake; 1999
Chi-Chi, Taiwan, earthquake; and 2008 Wenchuan, China, earthquake). In order to account for
uncertainties, the proposed logic tree incorporates mean () values and mean values 1 standard
deviation () for material properties (, c', |', and o
t
), slope geometry parameters (i and D
t
), and
the four horizontal and four vertical acceleration attenuation relations. When the proposed
factors were considered (vertical ground accelerations, topographic effects, and bond break
effects in addition to horizontal accelerations), the mean FS could capture the key features of LC
data (abrupt change and high concentration near fault).
A series of sensitivity analyses are also presented in this chapter. As the propose factors
were included, the standard deviations of FS were computed. For all of cases, standard deviation
decreases when horizontal acceleration was considered. Standard deviation also decreases as R
decreases. This is because of that ground accelerations have more influence on the FS as ground
accelerations become larger. In general, variation in slope angle results in the largest standard
deviation, whereas the variation of unit weight produces minor standard deviation.


128


Figure 8.1 Flow chart of new approach for the pseudo-static slope stability analysis. AS08, BA08,
CB08, and I 08 represent Abrahamson and Silva (2008), Boore and Atkinson (2008), and
Campbell and Bozorgnia (2008), and Idriss (2008), respectively. A 05, CB03, EP97, and AS96
represent Ambraseys et al. (2005), Campbell and Bozorgnia (2003), Elnashai and Papazouglou
(1997), and Ambraseys and Simpson (1996), respectively.

Step 1. Predict horizontal ground acceleration
using four attenuation relationships from NGA
project (AS08, BA08, CB08, and I08).
Step 5. Calculate factor of safety for infinite slope
using pseudo-static slope stability analysis including
horizontal and vertical accelerations.
Step 4. Check if soil bond breaks when the vertical
ground motion arrives.
Step 2. Predict vertical ground acceleration, using the
attenuation relationships by A 05 and CB03 and V/H
ratios proposed by EP97 and AS96.
Step 3. Apply topographic factor to both horizontal
and vertical accelerations.

129


Figure 8.2 Vertical acceleration along the distance from the fault predicted for the M
w
7.6
Kashmir, Pakistan, earthquake.

0 10 20 30 40 50
Distance from surface projection of updip edge of fault plane (km)
0
0.2
0.4
0.6
0.8
1
1.2
V
e
r
t
i
c
a
l

g
r
o
u
n
d

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Ambraseys et al. (2005)
Campbell-Bozorgnia (2003)
Elnashai-Papazouglou (1997)
Ambraseys-Simpson (1996)

130


Figure 8.3 Logic tree of slope model for the case of 2005 Kashmir, Pakistan, earthquake.
20
(0.2)
(0.6)
(0.2)
21
22
AS08

(kN/m
3
)
c'
(kPa)
|'
()
i
()
D
(m)
o
t

(kPa)
a
h
a
v T.F
k
v
k
h
(0.2)
(0.6)
(0.2)
(0.2)
(0.6)
(0.2)
(0.2)
(0.6)
(0.2)
(0.2)
(0.6)
(0.2)
(0.2)
(0.6)
(0.2)
BA08
CB08
I 08
19
27
35
34
40
46
0.58
0.73
0.88
2.33
3
3.67
18
28
38
(0.25)
(0.25)
(0.25)
(0.25)
(0.25)
(0.25)
(0.25)
(0.25)
(0.2)
(0.6)
(0.2)
(1/3)
(1/3)
1.3
1.45
1.6
0.9
0.33
Slope model
Material property Slope geometry Seismic parameter
A 05
CB03
EP97
AS96
as above
as above
as above as above
as above
as above
as above
as above
as below
as below as below
as below
(1/3)
0.5
(1/3)
(1/3)
0.9
0.33
(1/3)
0.5

131


Figure 8.4 Landslide concentration data for the 2005 Kashmir, Pakistan, earthquake (Sato et al.
2007) and factor of safety calculated by using various effects. The mean FS 1 standard
deviation are shown in a shaded area.
0 10 20 30 40 50
Distance from surface projection of updip edge of fault plane (km)
0
0.4
0.8
1.2
1.6
2
L
a
n
d
s
l
i
d
e

c
o
n
c
e
n
t
r
a
t
i
o
n
,

L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
2.4
2.6
2.8
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
LC for 2005 Kashmir, Pakistan earthquake
FS with k
h
+k
v
+Topographic effect + Bond break effect
FS with k
h
+k
v
+Topographic effect
FS with k
h
+ Topographic effect
FS with k
h
+k
v
FS with k
h
Static FS
Mean parameters
Tensile strength = 0.73 kPa
Unit weight = 21 kN/m
3
Cohesion intercept= 27 kPa
Friction angle = 40
o
Slope thickness = 3 m
Slope angle = 28
o
22

132


Figure 8.5 Mean factor of safety and mean 1 standard deviation predicted for the 2005 Kahsmir,
Pakistan, earthquake for adding each factor: (a) static condition, (b) with horizontal acceleration
(a
h
), (c) with a
h
and vertical acceleration (a
v
), (d) with a
h
and topographic effect, (e) with a
h
, a
v
,
and topographic effect, and (f) a
h
, a
v
, topographic effect, and bond break effect.
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
(a)
(b)
(c) (d)
(e) (f)

133


Figure 8.6 Results of sensitivity analysis for material properties and slope geometry parameters
for the 2005 Kahsmir, Pakistan, earthquake: (a) unit weight, (b) cohesion intercept, (c) friction
angle, (d) slope thickness, and (e) slope angle. Mean FS and mean FS 1 standard deviation ()
are shown for each parameter.

0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
FS with mean parameter
FS with mean parameter + 1 o
FS with mean parameter - 1 o
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
(a) (b)
(c)
(d)
(e)

134


Figure 8.7 Results of sensitivity analysis for (a) k
h
, (b) k
v
, (c) topographic factor, (d) horizontal
acceleration, and (e) vertical acceleration for the 2005 Kahsmir, Pakistan, earthquake.

0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
AS
BA
CB
I
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
A
CB
EP
AS
(a) (b)
(c) (d)
(e)
k
h
=0.33(a
h,max
/g)
0.5
0.5
T.F. =1.3
1.45
1.6
AS: Abrahamson-Silva 2008
BA: Boore-Atkinson 2008
CB: Campbell-Bozorgnia 2008
I: Idriss 2008
A: Ambraseys et al. 2005
CB: Campbell-Bozorgnia 2003
EP: Elnashai-Papazouglou 1997
AS: Ambraseys-Simpson 1996
0.7
k
v
=0.33(a
h,max
/g)
0.7

135


Figure 8.8 Mean factor of safety and mean 1 standard deviation by using the Monte-Carlo
simulation for (a) 2005 Kashmir, Pakistan earthquake, (b) 1989 Loma-Prieta earthquake, (d)
1999 Chi-Chi, Taiwan earthquake, and (d) 2008 Wenchuan, China earthquake.

136


Figure 8.9 Limits of southern Santa Cruz Mountains landslide area (left) and locations of earthquake-induced landslides with the
surface projection of fault rupture (right) (Keefer 2000)


137


Figure 8.10 Horizontal and vertical accelerations with respect to source-to-site distance predicted
for the M
w
6.9 1989 Loma-Prieta earthquake.
0 10 20 30 40 50
Distance from surface projection of updip edge of fault plane (km)
0
0.2
0.4
0.6
0.8
1
G
r
o
u
n
d

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Horizontal accelerations
Abrahamson and Silva (2008)
Boore and Atkinson (2008)
Campbell and Bozorgnia (2008)
Idriss (2008)
Vertical accelerations
Ambraseys et al. (2005)
Campbell and Bozorgnia (2003)
Elnashai and Papazouglou (1997)
Ambraseys and Simpson (1996)

138


Figure 8.11 Landslide concentration data for the 1989 Loma-Prieta earthquake by Keefer (2000)
and factor of safety calculated by using various effects. The mean FS 1 standard deviation are
shown in a shaded area.

0 10 20 30 40 50
Distance from surface projection of updip edge of fault plane (km)
0
0.4
0.8
1.2
1.6
2
L
a
n
d
s
l
i
d
e

c
o
n
c
e
n
t
r
a
t
i
o
n
,

L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
2.4
2.6
2.8
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
LC for 1989 Loma Prieta, CA earthquake
FS with k
h
+k
v
+Topographic effect + Bond break effect
FS with k
h
+k
v
+Topographic effect
FS with k
h
+ Topographic effect
FS with k
h
+k
v
FS with k
h
Static FS
Mean parameters
Tensile strength = 0.29 kPa
Unit weight = 21 kN/m
3
Cohesion intercept = 20 kPa
Friction angle = 33
o
Slope thickness = 3 m
Slope angle = 27
o

139


Figure 8.12 Mean factor of safety and mean 1 standard deviation predicted for the 1989 Loma-
Prieta, U.S., earthquake for adding each factor: (a) static condition, (b) with horizontal
acceleration (a
h
), (c) with a
h
and vertical acceleration (a
v
), (d) with a
h
and topographic effect, (e)
with a
h
, a
v
, and topographic effect, and (f) a
h
, a
v
, topographic effect, and bond break effect.

0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
(a)
(b)
(c) (d)
(e) (f)

140


Figure 8.13 Results of sensitivity analysis for material properties and slope geometry parameters
for the 1989 Loma-Prieta, U.S., earthquake: (a) unit weight, (b) cohesion intercept, (c) friction
angle, (d) slope thickness, and (e) slope angle. Mean FS and mean FS 1 standard deviation ()
are shown for each parameter.

0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
FS with mean parameter
FS with mean parameter + 1 o
FS with mean parameter - 1 o
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
(a)
(b)
(c) (d)
(e)

141


Figure 8.14 Results of sensitivity analysis for (a) k
h
, (b) k
v
, (c) topographic factor, (d) horizontal
acceleration, and (e) vertical acceleration for the 1989 Loma-Prieta, U.S., earthquake.
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
AS
BA
CB
I
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.4
0.8
1.2
1.6
2
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
A
CB
EP
AS
(a)
(b)
(c)
(d)
(e)
k
h
=0.33(a
h,max
/g)
0.5(a
h,max
/g)
k
v
=0.33(a
v,max
/g)
0.5(a
v,max
/g)
TF =1.30
1.45
1.60
AS: Abrahamson-Silva 2008
BA: Boore-Atkinson 2008
CB: Campbell-Bozorgnia 2008
I: Idriss 2008
A: Ambraseys et al. 2005
CB: Campbell-Bozorgnia 2003
EP: Elnashai-Papazouglou 1997
AS: Ambraseys-Simpson 1996
0.9(a
h,max
/g)
0.9(a
v,max
/g)

142


Figure 8.15 Locations of landslides induced by 1999 Chi-Chi, Taiwan, earthquake along the
Cher-Long-Pu fault and geological setting of affected area (Khazai and Sitar 2004)


143


Figure 8.16 Horizontal and vertical accelerations with respect to source-to-site distance predicted
for the M
w
7.6 1999 Chi-Chi, Taiwan, earthquake.
0 10 20 30 40 50
Distance from surface projection of updip edge of fault plane (km)
0
0.2
0.4
0.6
0.8
1
G
r
o
u
n
d

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Horizontal accelerations
Abrahamson and Silva (2008)
Boore and Atkinson (2008)
Campbell and Bozorgnia (2008)
Idriss (2008)
Vertical accelerations
Ambraseys et al. (2005)
Campbell and Bozorgnia (2003)
Elnashai and Papazouglou (1997)
Ambraseys and Simpson (1996)

144


Figure 8.17 Landslide concentration data for the 1999 Chi-chi, Taiwan, earthquake (Khazai and
Sitar 2004) and factor of safety calculated by using various effects. The mean FS 1 standard
deviation are shown in a shaded area.

0 10 20 30 40 50
Distance from surface projection of updip edge of fault plane (km)
0
0.2
0.4
0.6
0.8
1
L
a
n
d
s
l
i
d
e

c
o
n
c
e
n
t
r
a
t
i
o
n
,

L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
2.4
2.6
2.8
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
LC for 1999 Chi-Chi, Taiwan earthquake
FS with k
h
+k
v
+Topographic effect + Bond break effect
FS with k
h
+k
v
+Topographic effect
FS with k
h
+ Topographic effect
FS with k
h
+k
v
FS with k
h
Static FS
Mean parameters
Tensile strength = 0.73 kPa
Unit weight = 24 kN/m
3
Cohesion intercept = 11 kPa
Friction angle = 51
o
Slope thickness = 0.7 m
Slope angle = 65
o

145


Figure 8.18 Mean factor of safety and mean 1 standard deviation predicted for the 1999 Chi-
Chi, Taiwan, earthquake for adding each factor: (a) static condition, (b) with horizontal
acceleration (a
h
), (c) with a
h
and vertical acceleration (a
v
), (d) with a
h
and topographic effect, (e)
with a
h
, a
v
, and topographic effect, and (f) a
h
, a
v
, topographic effect, and bond break effect.


0
0.2
0.4
0.6
0.8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.2
0.4
0.6
0.8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.2
0.4
0.6
0.8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.2
0.4
0.6
0.8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.2
0.4
0.6
0.8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.2
0.4
0.6
0.8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
(a)
(b)
(c) (d)
(e) (f)

146


Figure 8.19 Results of sensitivity analysis for material properties and slope geometry parameters
for the 1999 Chi-Chi, Taiwan, earthquake: (a) unit weight, (b) cohesion intercept, (c) friction
angle, (d) slope thickness, and (e) slope angle. Mean FS and mean FS 1 standard deviation ()
are shown for each parameter.


0
0.2
0.4
0.6
0.8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.2
0.4
0.6
0.8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.2
0.4
0.6
0.8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.2
0.4
0.6
0.8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
FS with mean parameter
FS with mean parameter + 1 o
FS with mean parameter - 1 o
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.2
0.4
0.6
0.8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
(a) (b)
(c)
(d)
(e)

147


Figure 8.20 Results of sensitivity analysis for (a) k
h
, (b) k
v
, (c) topographic factor, (d) horizontal
acceleration, and (e) vertical acceleration for the 1999 Chi-Chi, Taiwan, earthquake.
0
0.2
0.4
0.6
0.8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.2
0.4
0.6
0.8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
0.2
0.4
0.6
0.8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.2
0.4
0.6
0.8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
AS
BA
CB
I
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
0.2
0.4
0.6
0.8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
A
CB
EP
AS
(a) (b)
(c) (d)
(e)
k
h
=0.33(a
h,max
/g)
0.5
T.F. =1.3
1.45
1.6
AS: Abrahamson-Silva 2008
BA: Boore-Atkinson 2008
CB: Campbell-Bozorgnia 2008
I: Idriss 2008
A: Ambraseys et al. 2005
CB: Campbell-Bozorgnia 2003
EP: Elnashai-Papazouglou 1997
AS: Ambraseys-Simpson 1996
0.9
k
v
=0.33(a
v,max
/g)
0.5
0.9

148


Figure 8.21 Distribution of earthquake-triggered landslides (black polygons) on the isoseismic map (Dai et al. 2011)

149


Figure 8.22 Horizontal and vertical accelerations with respect to source-to-site distance predicted
for the M
w
7.9 2008 Wenchuan, China, earthquake.
0 10 20 30 40 50
Distance from surface projection of updip edge of fault plane (km)
0
0.2
0.4
0.6
0.8
1
G
r
o
u
n
d

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Horizontal accelerations
Abrahamson and Silva (2008)
Boore and Atkinson (2008)
Campbell and Bozorgnia (2008)
Idriss (2008)
Vertical accelerations
Ambraseys et al. (2005)
Campbell and Bozorgnia (2003)
Elnashai and Papazouglou (1997)
Ambraseys and Simpson (1996)

150


Figure 8.23 Landslide concentration data for the 2008 Wenchuan, China, earthquake (Dai et al.
2010) and factor of safety calculated by using various effects. The mean FS 1 standard
deviation are shown in a shaded area.

0 10 20 30 40 50
Distance from surface projection of updip edge of fault plane (km)
0
1
2
3
4
5
6
7
8
L
a
n
d
s
l
i
d
e

c
o
n
c
e
n
t
r
a
t
i
o
n
,

L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
2.4
2.6
2.8
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
LC for 2008 Wenchuan, China earthquake
FS with k
h
+k
v
+ Topographic effect + Bond break effect
FS with k
h
+k
v
+ Topographic effect
FS with k
h
+ Topographic effect
FS with k
h
+k
v
FS with k
h
Static FS
Mean parameters
Tensile strength = 0.73 kPa
Unit weight = 24 kN/m
3
Cohesion intercept = 27 kPa
Friction angle = 40
o
Slope thickness = 3 m
Slope angle = 33
o

151


Figure 8.24 Mean factor of safety and mean 1 standard deviation predicted for the 2008
Wenchuan, China, earthquake for adding each factor: (a) static condition, (b) with horizontal
acceleration (a
h
), (c) with a
h
and vertical acceleration (a
v
), (d) with a
h
and topographic effect, (e)
with a
h
, a
v
, and topographic effect, and (f) a
h
, a
v
, topographic effect, and bond break effect.

0
2
4
6
8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
2
4
6
8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
2
4
6
8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
2
4
6
8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
2
4
6
8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
2
4
6
8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.5
1
1.5
2
2.5
3
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
(a)
(b)
(c) (d)
(e) (f)

152


Figure 8.25 Results of sensitivity analysis for material properties and slope geometry parameters
for the 2008 Wenchuan, China, earthquake: (a) unit weight, (b) cohesion intercept, (c) friction
angle, (d) slope thickness, and (e) slope angle. Mean FS and mean FS 1 standard deviation ()
are shown for each parameter.

0
2
4
6
8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
2
4
6
8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
2
4
6
8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
2
4
6
8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
FS with mean parameter
FS with mean parameter + 1 o
FS with mean parameter - 1 o
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
2
4
6
8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
(a) (b)
(c)
(d)
(e)

153


Figure 8.26 Results of sensitivity analysis for (a) k
h
, (b) k
v
, (c) topographic factor, (d) horizontal
acceleration, and (e) vertical acceleration for the 2008 Wenchuan, China, earthquake.

0
2
4
6
8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
2
4
6
8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0
2
4
6
8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
2
4
6
8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
AS
BA
CB
I
0 10 20 30 40 50
Distance from surface projection
of updip edge of fault plane (km)
0
2
4
6
8
L
C

(
l
a
n
d
s
l
i
d
e
s
/
k
m
2
)
0
0.4
0.8
1.2
1.6
2
2.4
F
a
c
t
o
r

o
f

s
a
f
e
t
y

a
g
a
i
n
s
t

s
l
i
d
i
n
g
A
CB
EP
AS
(a) (b)
(c) (d)
(e)
k
h
=0.33(a
h,max
/g)
0.5
T.F. =1.3
1.6
1.45
AS: Abrahamson-Silva 2008
BA: Boore-Atkinson 2008
CB: Campbell-Bozorgnia 2008
I: Idriss 2008
A: Ambraseys et al. 2005
CB: Campbell-Bozorgnia 2003
EP: Elnashai-Papazouglou 1997
AS: Ambraseys-Simpson 1996
0.9
k
v
=0.33(a
v,max
/g)
0.5
0.9

154

CHAPTER 9. CONCLUSIONS AND RECOMMENDATION FOR FUTURE RESEARCH
9.1 Summary and conclusions
Several seismic hazard studies have been recently conducted for regions of Pakistan
because of its active tectonic setting. However, the prior seismic studies used the diffuse areal
source zones, resulting in underestimation of seismic hazard near faults. The earthquake-induced
landslide is one of the most common secondary effects of earthquake as extensive landslides
occurred in NW Pakistan during the 2005 Kashmir earthquake. However, the currently used
approach of seismic slope stability analysis cannot capture the regional distribution of landslides.
The goal of this research is to assess the seismic hazard in NW Pakistan using discrete faults
rather than areal sources, and to propose the factors that affect the seismic slope stability and
explain the regional distribution of earthquake-induced landslides.
9.1.1 Seismic hazard analyses
The methodology of seismic hazard analysis in NW Pakistan was proposed in this
dissertation. In lieu of using the area source zones proposed by other studies (NORSAR and
PMD 2006; PMD and NORSAR 2006; Monalisa et al. 2007), 32 discrete faults were employed
with their surface traces, geometries, and faulting mechanisms. Then the earthquake catalog was
used to construct the exponential recurrence models. For activity rate of the exponential
recurrence model, two different period of earthquake catalogs were used; including and
excluding the earthquake event in 2005. This was because of that extremely high amount of
foreshocks and aftershocks of 2005 Kashmir earthquake could bias the result. The earthquake
data from the recent 30-year interval were used to estimate the activity rate and b-value because
earthquakes for 4M
w
<6 are considered to be incomplete prior to 1975. The large magnitude
range of computed recurrence model matches with the earthquake data for 100-year interval. In
order to estimate characteristic rate of the characteristic recurrence model, three methods were
proposed. The first method assumes that the characteristic rate is related with the slip rate of
faults. The second method employs the assumption that the characteristic rate is proportional to
the activity rate. The third method is using the seismic moment balance method proposed by Aki
(1966). All these parameters and models were implemented in the logic tree and reasonable
weight for each branch of the logic tree was assigned. Finally four NGAs were selected based on

155

the fact that they can capture the measured ground acceleration during the 2005 Kashmir,
Pakistan, earthquake, and were used according to the fault models developed in this study.
Seismic hazard curves and seismic hazard spectra were generated for selected cities.
Seismic hazard maps corresponding to 475-, 975-, and 2475-year return periods were also
generated for the region of NW Pakistan. The results show that cities with the highest hazards
are Kaghan and Muzaffarabad, with PGA ~ 0.6g for a 475-year return period. Islamabad, the
capital city of Pakistan, also has a significant seismic hazard, much higher than predicted by
previous studies. On the other hand, much lower ground motions were predicted for the cities
located farther from active faults, including Astor, Bannu, Malakand, Mangla, Peshawar, and
Talagang. However, the PGAs computed for these cities are still similar to or slightly higher than
those by previous studies.
Despite the uncertainties and assumptions associated with both the deterministic and
probabilistic seismic hazard analyses, the results from both methodologies were comparable,
thus providing a measure of confidence in the results. The computed hazards also correspond
well to the known seismic history of the region, including the recent 2005 Kashmir earthquake.
In addition, use of the conditional mean spectrum instead of the uniform hazard spectrum could
provide significant efficiency in seismic design in many cities in NW Pakistan.
9.1.2 Seismic slope stability analyses
When earthquakes occur in mountainous regions, earthquake-induced landslides and
rockfalls commonly cause significant damage. However, current practice of using pseudo-static
slope stability analysis to evaluate landslides only accounts for the horizontal ground
acceleration, and does not predict factors of safety that are consistent with available landslide
concentration data. This study proposed three other factors that can influence rock slope stability:
(1) vertical ground accelerations; (2) topographic effects; and (3) bond break effects.
The vertical ground acceleration, when estimated using available attenuation relations
and vertical-to-horizontal acceleration ratios, can equal the horizontal acceleration in the near-
fault region for large magnitude earthquakes. For gentle to moderate slopes (i.e., less than ~ 60)
and sliding depths exceeding about 1 m, incorporating large vertical accelerations in a pseudo-
static stability analysis tends to decrease the FS. While the effect of vertical acceleration on the
computed FS is not significant, this study suggests that it should not be ignored. Furthermore,

156

both the horizontal and vertical accelerations may be amplified near slope crests as a result of
constructive geometric wave interference, i.e., topographic effects. When the vertical
acceleration was amplified by a topographic factor, the effect of vertical accelerations on FS
became more important. Lastly, the occurrence of high amplitude vertical accelerations on slopes
may damage or break interparticle bonding (i.e., cementation, cohesion, etc.) along rock
bedding planes, particularly near the ground surface where weathering and jointing is commonly
more severe, effectively eliminating the equivalent cohesion of the rock. If bonding is destroyed,
the FS can decrease considerably. In fact, observations in many earthquakes suggest that most
earthquake-induced landslides are shallow, disrupted slides and falls that occur near slope crests
or peaks, where topographic amplification and bond break effects are most likely to occur.
In this dissertation, a new approach to evaluate regional pseudo-static slope stability that
accounts for vertical accelerations, potential topographic amplification, and potential bond
breakage was proposed. The procedure was incorporated into a logic tree approach to account for
variability in both rock and geometric properties of the slope. To examine the applicability of the
newly proposed method, factors of safety computed with this procedure was compared to
landslide concentration data collected during four recent earthquakes (1989 Loma-Prieta, 1999
Chi-Chi, 2005 Kashmir, and 2008 Wenchuan). Observations from these events all show abrupt
changes in landslide concentration in the near-fault region. When the computed FS incorporates
vertical accelerations, topographic effects, and bond breakage, the FS trends are consistent with
the observed trends in landslide concentration, and qualitatively explain the abrupt increase in
landslide concentration in the near-fault region of these earthquakes.
9.2 Recommendations for future work
1) Although this study estimated some of the input parameters based on best estimates
due to lack of information on fault properties in NW Pakistan, it is necessary to obtain
these parameters as appropriately as possible in the future to yield the most reliable
results. Obtaining the precise fault width and dip is very important task in seismic
hazard analysis not only because the fault geometry determines the distance between
source and site, but also because it affects the determination of upper bound
magnitude. Slip rate of individual fault is another important fault property that needs

157

to be well investigated because it plays a significant role in estimating the proper
characteristic rate.
2) Unfortunately, no region-specific ground motion prediction model is available for
Pakistan although this region is seismically highly active. Developing the GMPE that
is suitable for local conditions in Pakistan is the essential task that will improve the
seismic hazard assessment of this region.
3) There are seismic hazard studies conducted based on the area source zone for various
regions because the faults are not well-characterized, resulting in underestimation of
seismic hazard. It is necessary to assess the seismic hazard using available
information on discrete faults for these regions.
4) The proposed factors that affect the slope stability need to be applied to other
methods including Newmarks sliding method and energy-based method. The new
approach proposed in this thesis can also be extended to evaluate the stability of
individual slopes.






158


APPENDICES


159

APPENDIX A. ESTIMATION OF MOHR-COULOMB STRENGTH
PARAMETERS
A1. DETERMINATION OF PARAMETERS FOR HOEK-BROWN CRITERION
1) Uniaxial compressive strength of the intact rock,
ci

No information is available for the uniaxial compressive strength of intack rock for the
failure mass of four earthquake cases in this study. Therefore, o
ci
= 2 MPa was considered based
on the typical value for highly weathered rock (Table A. 1) suggested by Hoek et al. (1998) for
the case of 1999 Chi-Chi, 2005 Kahsmir, and 2008 Wenchuan earthquakes. Measured values of
friction angle and cohesion intercept are available for the case of 1989 Loma-Prieta earthquake
by McCrink and Real (1996). The o
ci
of 0.8 MPa was then estimated to yield the equivalent
Mohr-Coulomb parameters consistent with the measured mean friction angle and median
cohesion intercept of 33.1 and 25.2 kPa, respectively.

2) Material constant for intact rock, m
i

m
i
can be obtained from Table A. 2. Majority of landslides occurred in sedimentary rocks
which consist of sandstone and siltstone for all cases of earthquake. Therefore the average of mi
for sandstone (19) and siltstone (9) was used in the analysis.

3) Geological strength index, GSI
GSI classification is based on an assessment of the lithology, structure and condition of
discontinuity surfaces in the rock mass (Marinos et al. 2005). Based on the description of failure
mass, it can be concluded for all earthquake cases that the surface conditions of failure masses
are highly weathered, and the structures are blocky/disturbed and folded. Therefore, the GSI of
30 was obtained from Figure A. 1.

The descriptions for surface condition and structure of failure mass for four earthquake
cases are as below:


160

2005 Kashmir, Pakistan, earthquake: Kamp et al. (2008) described the failed sedimentary rocks
as undeformed to tightly-folded, highly-cleaved and fractured. Owen et al. (2008) observed that
these shallow landslides typically involved the top few meters of weathered bedrock, regolith,
and soil.

1989 Loma-Prieta, U.S., earthquake: Keefer (2000) reported that the Purisima Formation in
which most of landslides occurred consists of weakly cemented sandstones and siltstones, and
the rocks are poorly to moderately indurated, and structurally deformed by pervasive folding and
local faulting. Keefer and Manson (1989) also reported that disrupted landslides in all units
occurred in materials that typically were weakly cemented, closely fractured, intensely
weathered, and broken by joints.

1999 Chi-Chi, Taiwan, earthquake: The landslide masses commonly were dry, highly
disaggregated, weakly cemented, closely fractured, intensely weathered, and broken by
conspicuous, highly persistent joints (Khazai and Sitar 2004).

2008 Wenchuan, China, eartqhauke: Yin et al. (2009) and Dai et al. (2011) reported that most of
the landslides occurred in weathered or heavily fractured sandstone, siltstone, slate, and phyllite.

4) Disturbance factor (DF)
Hoek et al. (2002) proposed the use of disturbance factor (DF) to account for the effects
of heavy blast damage and stress relief due to removal of the overburden. As shown in Table A.
3, DF = 0 is suggested when there is no damage from blasting or excavation, whereas DF= 1 is
suggested for the rocks affected by heavy blast damage or stress relief. The DF was considered
to be 0 in this study because there was no evidence for disturbance due to heavy blasting and
stress relief for all earthquake cases.



161

A2. ESTIMATION OF TENSILE STRENGTH AND MOHR-COULOMB STRENGTH
PARAMETERS
Hoek and Brown (Hoek and Brown 1980) proposed the failure criterion for heavily
jointed rock masses which can be expressed as:

5 . 0
3
3 1
|
|
.
|

\
|
+
'
+ ' = ' s m
ci
ci
o
o
o o o Eq. (A-1)

where
1
o'
and
3
o' are the major and minor effective principal stresses
ci
o is the uniaxial compressive strength of the intact rock
m and s are material constants, where s =1 for intact rock.

The generalized Hoek-Brown criterion takes into account the GSI, and can be expressed
as (Hoek et al. 2002):

a
ci
b ci
s m
|
|
.
|

\
|
+
'
+ ' = '
o
o
o o o
3
3 1
Eq. (A-2)

where m
b
is a reduced value of the material constant m
i,
and is given by

|
.
|

\
|


=
DF
GSI
m m
i b
14 28
100
exp
Eq. (A-3)

|
.
|

\
|


=
DF
GSI
s
3 9
100
exp
Eq. (A-4)

( )
3 20 15
6
1
2
1

+ = e e a
GSI
Eq. (A-5)

The tensile strength can be computes as:

162

b
ci
t
m
so
o =
Eq. (A-6)

Since most of slope stability analysis requires the Mohr-Coulomb strength parameters
that are not directly measured for rock, it is necessary to determine equivalent friction angle and
cohesion intercept parameters for failure mass. This can be done by fitting an average linear
relationship to the Hoek-Brown criterion for an appropriate stress range.

Using the fitting process that involves balancing the areas above and below the Mohr-
Coulomb plot, Hoek et al. (2002) proposed the following equations for |' and c':

( )
( )( ) ( ) ( )
(
(

' + + + +
' +
= '

1
3
1
3 1
6 1 2 1 2
6
sin
a
n b b
a
n b b
m s am a a
m s am
o
o
| Eq. (A-7)

( ) ( ) | |( )
( )( )
( )( )
( )( ) a a
m s am
a a
m s m a s a
c
a
n b b
a
n b n b ci
+ +
' + +
+ +
' + ' + +
= '

2 1
6 1
2 1
1 2 1
1
3
1
3 3
o
o o o


where
ci
n
o
o
o
max 3
3
'
= '
Eq. (A-8)

Figure A. 2 shows Hoek-Brown criteria and Mohr-Coulomb fits corresponding to
appropriate stress ranges based on thickness and unit weight of failure mass (i.e. approximately
0.06-0.08 MPa for 2005 Kahsmir, 1989 Loma-Prieta, and 2008 Wenchuan earthquakes, and
approximately 0.02 MPa for 1999 Chi-Chi earthquake).

163

Table A. 1 Field estimates of the uniaxial compressive strength of intact pieces (Hoek et al. 1998).
Term
Uniaxial
compressive
strength (MPa)
Point load
index
(MPa)
Field estimate of strength Examples
Extremely
strong
>250 >10
Specimen can only be chipped with a
geological hammer
Fresh basalt, chert, diabase, gneiss,
granite, quartzite
Very
strong
100-250 4-10
Specimen requires many blows of a
geological hammer to fracture it
Amphibolite, sandstone, basalt, gabbro,
gneiss, granodiorite, limestone, marble,
rhyolite, tuff
Strong 50-100 2-4
Specimen requires more than one blow of a
geological hammer to fracture it
Limestone, marble, phyllite, sandstone,
schist, shale
Medium
strong
25-50 1-2
Cannot be scraped or peeled with a pocket
knife, specimen can be fractured with a
single blow from a geological hammer
Claystone, coal, concrete, schist, shale,
siltstone
Weak 5-25 *
Can be peeled with a pocket knife with
difficulty, shallow indentation made by firm
blow with point of a geological hammer
Chalk, rocksalt, potash
Very weak 1-5 *
Crumbles under firm blows with point of a
geological hammer, can be peeled by a
pocket knife
Highly weathered or altered rock
Extremely
weak
0.25-1 * Indented by thumbnail Stiff fault gouge
* Point load tests on rocks with a uniaxial compressive strength below 25 MPa are likely to yield ambiguous results

164

Table A. 2 Values of constant mi for intact rock. Values in parenthesis are estimates (Hoek et al. 1998).
Rock type Class Group
Texture
Coarse Medium Fine Very fine
Sedimentary Clastic
Conglomerate
(22)
Sandstone
19

Siltstone
9
Claystone
4

Greywacke
(18)

Non-clastic Organic
Chalk
7


Coal
(8-21)

Carbonate
Breccia
(20)
Sparitic
limestone
(10)

Micritic
limestone
8

Chemical
Gypstone
16

Anhydrite
13

Metamorphic Non-foliated
Marble
9
Hornfels
19

Quartzite
24

Slightly foliated
Migmatite
(30)
Amphilbolite
25-31

Mylonites
(6)

Foliated
Gneiss
33
Schists
4-8

Phyllites
(10)
Slate
9
Igneous Light
Granite
33

Rhyolite
(16)
Obsidian
(19)

Granodiorite
(30)

Dacite
(17)


Diorite
(28)

Andesite
19

Dark
Gabbro
27
Dolerite
(19)

Basalt
(17)


Norite
22


Extrusive
pyroclastic type

Agglomerate
(20)
Breccia
(19)

Tuff
(15)



165

Table A. 3 Guidelines for estimating disturbance factor, DF (Hoek et al. 2002).
appearance of rock mass Description of rock mass
Suggested
value of DF

Excellent quality controlled blasting or
excavation by Tunnel Boring Machine results in
minimal disturbance to the confined rock mass
surrounding a tunnel.
DF = 0

Mechanical or hand excavation in poor quality
rock masses (no blasting) results in minimal
disturbance to the surrounding rock mass.
DF = 0
Where squeezing problems result in significant
floor heave, disturbance can be severe unless a
temporary invert, as shown in the photograph, is
placed.
DF = 0.5
No invert

Very poor quality blasting in a hard rock tunnel
results in severe local damage, extending 2 or 3
m, in the surrounding rock mass.
DF = 0.8

Small scale blasting in civil engineering slopes
results in modest rock mass damage, particularly
if controlled blasting is used as shown on the left
hand side of the photograph. However, stress
relief results in some disturbance.
DF = 0.7
Good blasting

DF = 1.0
Poor blasting

Very large open pit mine slopes suffer significant
disturbance due to heavy production blasting and
also due to stress relief from overburden
removal.
In some softer rocks excavation can be carried
out by ripping and dozing and the degree of
damage to the slopes is less.
DF = 1.0
Production
blasting

DF = 0.7
Mechanical
excavation



166


Figure A. 1 Geological strength index for jointed rocks (Hoek and Marinos 2000).



167


Figure A. 2 Hoek-Brown criteria and corresponding Mohr-Coulomb fits for (a) 2005 Kahsmir,
Pakistan, earthquake, (b) 1989 Loma-Prieta, U.S. earthquake, (c) 1999 Chi-Chi, Taiwan,
earthquake, and (d) 2008 Wenchuan, China, earthquake.


0
0.04
0.08
0.12
0.16
0.2
S
h
e
a
r

s
t
r
e
s
s

(
M
P
a
)
Hoek-Brown Criterion
Mohr-Coulomb Fit
0 0.05 0.1 0.15 0.2 0.25
Effective normal stress (MPa)
0
0.04
0.08
0.12
0.16
0.2
S
h
e
a
r

s
t
r
e
s
s

(
M
P
a
)
0 0.05 0.1 0.15 0.2 0.25
Effective normal stress (MPa)
(a) (b)
(c) (d)

168

REFERENCES
Abrahamson, N., Atkinson, G., Boore, D., Bozorgnia, Y., Campbell, K., Chiou, B., Idriss, I. M.,
Silva, W. and Youngs, R. (2008). "Comparisons of the NGA ground-motion relations."
Earthquake Spectra 24(1): 45-66.
Abrahamson, N. and Silva, W. (2008). "Summary of the Abrahamson & Silva NGA ground-
motion relations." Earthquake Spectra 24(1): 67-97.
Abrahamson, N. A. and Litehiser, J. J. (1989). "Attenuation of vertical peak acceleration."
Bulletin - Seismological Society of America 79(3): 549-580.
Aki, K. (1966). "Generation and propagation of G-waves from the Niigata earthquake of June
19, 1964. Part 2. Estimation of earthquake movement, released energy, and stress-strain
drop from G-wave spectrum." Bull. Earthquake Res. Inst. 44: 23-88.
Aldama-Bustos, G., Bommer, J. J., Fenton, C. H. and Stafford, P. J. (2009). "Probabilistic
seismic hazard analysis for rock sites in the cities of Abu Dhabi, Dubai and Ra's Al
Khaymah, United Arab Emirates." Georisk 3(1): 1-29.
Ambraseys, N. and Bilham, R. (2000). "A note on the Kangra Ms = 7.8 earthquake of 4 April
1905." Current Science 79(1): 45-50.
Ambraseys, N. N. and Douglas, J. (2004). "Magnitude calibration of north Indian earthquakes."
Geophysical Journal International 159(1): 165-206.
Ambraseys, N. N., Douglas, J., Sarma, S. K. and Smit, P. M. (2005a). "Equations for the
estimation of strong ground motions from shallow crustal earthquakes using data from
Europe and the middle east: Horizontal peak ground acceleration and spectral
acceleration." Bulletin of Earthquake Engineering 3(1): 1-53.
Ambraseys, N. N., Douglas, J., Sarma, S. K. and Smit, P. M. (2005b). "Equations for the
estimation of strong ground motions from shallow crustal eUsing data from Europe and
the middle east: Vertical peak ground acceleration and spectral acceleration." Bulletin of
Earthquake Engineering 3(1): 55-73.
Ambraseys, N. N. and Simpson, K. A. (1995). Prediction of vertical response spectra in Europe.
Research Report ESEE-95/1. Imperial College.
Ambraseys, N. N. and Simpson, K. A. (1996). "Prediction of vertical response spectra in
Europe." Earthquake Engineering and Structural Dynamics 25(4): 401-412.
Ambraseys, N. N., Simpson, K. A. and Bommer, J. J. (1996). "Prediction of horizontal response
spectra in Europe." Earthquake Engineering and Structural Dynamics 25(4): 371-400.
Ashford, S. A. and Sitar, N. (1997). "Analysis of topographic amplification of inclined shear
waves in a steep coastal bluff." Bulletin of the Seismological Society of America 87(3):
692-700.
Ashford, S. A. and Sitar, N. (2002). "Simplified method for evaluating seismic stability of steep
slopes." Journal of Geotechnical and Geoenvironmental Engineering 128(2): 119-128.
Ashford, S. A., Sitar, N., Lysmer, J. and Deng, N. (1997). "Topographic effects on the seismic
response of steep slopes." Bulletin of the Seismological Society of America 87(3): 701-
709.
Assimaki, D. and Gazetas, G. (2004). "Soil and topographic amplification on canyon banks and
the 1999 Athens earthquake." Journal of Earthquake Engineering 8(1): 1-43.
Assimaki, D., Kausel, E. and Gazetas, G. (2005). "Soil-dependent topographic effects: A case
study from the 1999 Athens earthquake." Earthquake Spectra 21(4): 929-966.

169

Aydan, ., Ohta, Y. and Hamada, M. (2009). "Geotechnical evaluation of slope and ground
failures during the 8 October 2005 Muzaffarabad earthquake, Pakistan." Journal of
Seismology 13(3): 399-413.
Baker, J. W. (2008). "The conditional mean spectrum: A tool for ground motion selection."
Journal of structural engineering in review.
Baker, J. W. and Cornell, C. A. (2006). "Spectral shape, epsilon and record selection."
Earthquake Engineering and Structural Dynamics 35(9): 1077-1095.
Baker, J. W. and Jayaram, N. (2008). "Correlation of spectral acceleration values from NGA
ground motinos models." Earthquake Spectra 24(1): 299-317.
Bazzurro, P. and Cornell, C. A. (1999). "Disaggregation of seismic hazard." Bulletin of the
Seismological Society of America 89(2): 501-520.
Bendick, R., Bilham, R., Khan, M. A. and Khan, S. F. (2007). "Slip on an active wedge thrust
from geodetic observations of the 8 October 2005 Kashmir earthquake." Geology 35(3):
267-270.
Bernreuter, D. L. (1992). "Determining the Controlling Earthquake from Probabilistic Hazards
for the Proposed Appendix B." Lawrence Livermore National Laboratory Report UCRL-
JC-111964.
Bhat, G. M., Pandita, S. K., Singh, Y., Sharma, V., Singh, S. and Bhat, G. R. (2005). October 8
Kashmir Earthquake: Impact on geoenvironment and structures in the Karnah and Uri
Tehsils of Kashmir (India). Post Graduate Department of Geology, Univeristy of Jammu.
Bilham, R. (2004). "Historical Studies on Earthqaueks in India." Annals of Geophysics.
Bilham, R. and Ambraseys, N. (2005). "Apparent Himalayan slip deficit from the summation of
seismic moments for Himalayan earthquakes, 1500-2000." Current Science 88(10): 1658-
1663.
Bonilla, M. G., Mark, R. K. and Lienkaemper, J. J. (1984). "Statistical relations among
earthquake magnitude, surface rupture length, and surface fault displacement." Bulletin
of the Seismological Society of America 74(6): 2379-2411.
Boore, D. M. (1972). "Note on effect of simple topography on seismic SH waves." Bulletin of
the Seismological Society of America 62(1): 275-&.
Boore, D. M. and 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 24(1): 99-138.
Boore, D. M., Joyner, W. B. and 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.
Bouckovalas, G. D. and Papadimitriou, A. G. (2005). "Numerical evaluation of slope topography
effects on seismic ground motion." Soil Dynamics and Earthquake Engineering 25(7-10):
547-558.
Bozzoni, F., Corigliano, M., Lai, C. G., Salazar, W., Scandella, L., Zuccolo, E., Latchman, J.,
Lynch, L. and Robertson, R. (2011). "Probabilistic seismic hazard assessment at the
eastern Caribbean Islands." Bulletin of the seismological society of America 101(5):
2499-2521.
Bray, J. D., Rathie, E. M., Augello, A. J. and Merry, S. M. (1998). "Simplified seismic design
procedure for geosynthetic-lined, solid-waste landfills." Geosynthetics International 5(1-
2): 203-235.

170

Brown, P. T. and Booker, J. R. (1985). "Finite Element Analysis of Excavation." Computers and
Geotechnics 1: 207-220.
Campbell, K. W. and Bozorgnia, Y. (2003). "Updated near-source ground-motion (attenuation)
relations for the horizontal and vertical components of peak ground acceleration and
acceleration response spectra." Bulletin of the Seismological Society of America 93(1):
314-331.
Campbell, K. W. and Bozorgnia, Y. (2007). Campbell-Bozorgnia NGA ground motion relations
for the geometric mean horizontal component of peak and spectral ground motion
parameters. Pacific Earthquake Engineerign Research Center.
Campbell, K. W. and 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.
Castellani, A., Chesi, C., Peano, A. and Sardella, L. (1982). "Seismic response of topographic
irregularities." Soil dynamics and earthquake engineering. Proc. conference,
Southampton, July 1982. Vol. 1: 251-268.
Celebi, M. (1987). "Topographical and geological amplifications determined from strong-motion
and aftershock records of the 3 March 1985 Chile earthquake." Bulletin of the
Seismological Society of America 77(4): 1147-1167.
Celebi, M. (1991). "Topographical and geological amplification: case studies and engineering
implications." Structural Safety 10(1-3): 199-217.
Chiou, B. S. J. and Youngs, R. R. (2008). "An NGA model for the average horizontal component
of peak ground motion and response spectra." Earthquake Spectra 24(1): 173-215.
Collier, C. J. and Elnashai, A. S. (2001). "A procedure for combining vertical and horizontal
seismic action effects." Journal of Earthquake Engineering 5(4): 521-539.
Cornell, C. A. and Vanmarcke, E. H. (1969). The major influences on seismic risk. Proceedings
of the Fourth World Conferences on Earthquake Engineering. Santiago, Chile. A(1): 69-
93.
Cui, P., Zhu, Y. Y., Han, Y. S., Chen, X. Q. and Zhuang, J. Q. (2009). "The 12 May Wenchuan
earthquake-induced landslide lakes: Distribution and preliminary risk evaluation."
Landslides 6(3): 209-223.
Dai, F. C., Xu, C., Yao, X., Xu, L., Tu, X. B. and Gong, Q. M. (2011). "Spatial distribution of
landslides triggered by the 2008 Ms 8.0 Wenchuan earthquake, China." Journal of Asian
Earth Sciences 40(4): 883-895.
Dai, S.-H. and Wang, M.-O. (1992). Reliability analysis in Engineering Applications. Van
Nostrand Reinhold. New York.
Dasgupta, S., Pande, P., Ganguly, D., Iqbal, Z., Sanyal, K., Venkatraman, N. V., Dasgupta, S.,
Sural, B., Harendranath, L., Mazumdar, K., Sanyal, S., Roy, A., Das, L. K., Misra, P. S.
and Gupta, H. (2000). Seismotectonic atlas of India and its environs. P. L. Narula, S. K.
Acharyya and J. Banerjee. Geological Survey of India.
Dipietro, J. A., Hussain, A., Ahmad, I. and Khan, M. A. (2000). "The main mantle thrust in
Pakistan: Its character and extent." Geological Society Special Publication: 375-393.
Dunning, S. A., Mitchell, W. A., Rosser, N. J. and Petley, D. N. (2007). "The Hattian Bala rock
avalanche and associated landslides triggered by the Kashmir Earthquake of 8 October
2005." Engineering Geology 93(3-4): 130-144.

171

Durrani, A. J., Elnashai, A. S., Hashash, Y. M. A., Kim, S.-J. and Masud, A. (2005). The
Kashmir Earthquake of October 8, 2005. A quick look report. Mid-America Earthquake
Center, University of Illinois at Urbana-Champaign. Urbana: 51 p.
EC8 (2000). Design Provisions for Earthquake Resistance of Structures, Prt 1-1: General rules-
Seismic Actions and General Requirements for Structures. prEN. 1998-5.
Elnashai, A. S. and Papazoglou, A. J. (1997). "Procedure and spectra for analysis of RC
structures subjected to strong vertical earthquake loads." Journal of Earthquake
Engineering 1(1): 121-155.
French Association for Earthquake Engineering (AFPS) (1995). Guidelines for Seismic
Microzonation studies. October.
Gardner, J. K. and Knopoff, L. (1974). "Is the sequence of earthquakes in Southern California,
with aftershocks removed, Poissonian?" Bulletin of the seismological society of America
64(5): 1363-1367.
Gazetas, G., Kallou, P. V. and Psarropoulos, P. N. (2002). "Topography and soil effects in the
Ms 5.9 Parnitha (Athens) earthquakes: The case of Admes." Natural Hazards 27(1-2):
133-169.
Graizer, V. (2009). "Low-velocity zone and topography as a source of site amplification effect
on Tarzana hill, California." Soil Dynamics and Earthquake Engineering 29(2): 324-332.
Grunthal, G. and Wahlstrom, R. (2003). "An Mw based earthquake catalogue for central,
northern and northwestern Europe using a hierarchy of magnitude conversions." Journal
of Seismology 7.
Gutenberg, B. and Richter, C. F. (1954). Seismicity of the earth and associated phenomena.
Princeton University Press. Princeton, N.J.
Hanks, T. C. and Bakun, W. H. (2002). "A bilinear source-scaling model for M-log A
observations of continental earthquakes." Bulletin of the Seismological Society of
America 92(5): 1847-1846.
Hanks, T. C. and Bakun, W. H. (2008). "M-log A observations for recent large earthquakes."
Bulletin of the Seismological Society of America 98(1): 490-494.
Hanks, T. C. and Kanamori, H. (1979). "A moment magnitude scale." Journal of Geophysical
Research B: Solid Earth 84(B5): 2348-2350.
Hartzell, S. H., Carver, D. L. and King, K. W. (1994). "Initial investigation of site and
topographic effects at Robinwood Ridge, California." Bulletin - Seismological Society of
America 84(5): 1336-1349.
Heuckroth, L. E. and Karim, R. A. (1973). "AFGHAN SEISMOTECTONICS." Philosophical
Transactions of the Royal Society of London Series a-Mathematical Physical and
Engineering Sciences 274(1239): 389-395.
Hoek, E. and Bray, J. (1981). Rock slope engineering. Institution of Mining and Metallurgy.
London.
Hoek, E. and Brown, E. T. (1980). Underground Excavations in Rock. Institution of Mining and
Metallurgy. London.
Hoek, E., Carranza-Torres, C. and Corkum, B. (2002). "Hoek-Brown failure criterion-2002
edition." Proceedings of the Fifth North American Rock Mechanics Symposium 1: 267-
273.
Hoek, E., Marinos, P. and Benissi, M. (1998). "Applicability of the geological strength index
(GSI) classification for very weak and sheared rock masses. The case of the Athens

172

Schist Formation." Bulletin of Engineering Geology and the Environment 57(2): 151-
160.
Hough, S. E., Altidor, J. R., Anglade, D., Given, D., Janvier, M. G., Maharrey, J. Z., Meremonte,
M., Mildor, B. S. L., Prepetit, C. and Yong, A. (2010). "Localized damage caused by
topographic amplification during the 2010 M7.0 Haiti earthquake." Nature Geoscience
3(11): 778-782.
Hung, J. J. (2000). "Chi-Chi earthquake induced landslides in Taiwan." Earthquake Engineering
and Engineering Seismology 2(2): 25-33.
Hynes-Griffin, M. E. and Franklin, A. G. (1984). Rationalizing the seismic coefficient method.
Miscellaneous Paper GL-84-13. U. S. Army Corps of Engineers Waterways Experiment
Station. Vicksburg, Mississippi: 521 pp.
Idriss, I. M. (2008). "An NGA empirical model for estimating the horizontal spectral values
generated by shallow crustal earthquakes." Earthquake spectra 24(1): 217-242.
Idriss, I. M. and Seed, H. B. (1967). "Response of Earth Banks During Earthquakes." Soil
Mechanics and Foundations Division, ASCE 93(SM3): 61-82.
Jaiswal, K. and Sinha, R. (2007). "Probabilistic seismic-hazard estimation for peninsular India."
Bulletin of the seismological society of America 97(1 B): 318-330.
Jayangondaperumal, R., Thakur, V. C. and Suresh, N. (2008). "Liquefaction features of the 2005
Muzaffarabad-Kashmir earthquake and evidence of palaeoearthquakes near Jammu,
Kashmir Himalaya." Current Science 95(8): 1071-1077.
Johnston, A. C. (1996). "Seismic moment assessment of earthquakes in stable continental
regions - I. Instrumental seismicity." Geophysical Journal International 124(2): 381-414.
Kamp, U., Growley, B. J., Khattak, G. A. and Owen, L. A. (2008). "GIS-based landslide
susceptibility mapping for the 2005 Kashmir earthquake region." Geomorphology
101(4): 631-642.
Kaul, P. H. (1911). Census of India.
Keefer, D. K. (1984). "Landslides caused by earthquakes." Geological Society of America
Bulletin 95(4): 406-421.
Keefer, D. K. (2000). Statistical analysis of an earthquake-induced landslide distribution - the
1989 Loma Prieta, California event: 231-249.
Keefer, D. K. (2002). "Investigating landslides caused by earthquakes - A historical review."
Surveys in Geophysics 23(6): 473-510.
Keefer, D. K. and Manson, M. W. (1989). Regional distribution and characteristics of landslides
generated by the earthquake. The Loma Prieta, California, earthquake of October 17,
1989: Strong ground motion and ground failure. U.S. Geological Survey Professional
Paper 1551-C.
Keefer, D. K. and Manson, M. W. (1998). "Regional distribution and characteristics of landslides
generated by the earthquake." US Geological Survey Professional Paper 1551-C: 7-32.
Keefer, D. K. and Wilson, R. C. (1989). "Predicting earthquake-induced landslides, with
emphasis on arid and semi-arid environments." Landslides in a Semi-Arid Environment
2(PART 1): 118-149.
Khazai, B. and Sitar, N. (2004). "Evaluation of factors controlling earthquake-induced landslides
caused by Chi-Chi earthquake and comparison with the Northridge and Loma Prieta
events." Engineering Geology 71(1-2): 79-95.

173

Kim, B., Hashash, Y. M. A., Olson, S. M. and Ahmad, I. (2010). Probabilistic seismic hazard
analysis for Islamabad and Peshawar in Pakistan using discrete faults. 5th International
Conference on Earthquake Geotechnical Engineering. Santiago, Chile.
Kokusho, T. and Ishizawa, T. (2007). "Energy approach to earthquake-induced slope failures and
its implications." Journal of Geotechnical and Geoenvironmental Engineering 133(7):
828-840.
Kokusho, T. and Kabasawa, K. (2003). "Energy approach to flow failure and its application to
flow due to water film in liquefied deposits." Proc., Int. Conf. on Fast Slope Movement:
297-302.
Kovacs, W. D., Seed, H. B. and Idriss, I. M. (1971). "Studies of seismic response of clay banks."
ASCE J Soil Mech Found Div 97(SM2).
Kramer, S. L. (1996). Geotechnical earthquake engineering. Prentice Hall. Upper Saddle River,
N.J.: xviii, 653.
Kramer, S. L. and Smith, M. W. (1997). "Modified Newmark model for seismic displacements
of compliant slopes." Journal of Geotechnical and Geoenvironmental Engineering
123(7): 635-644.
Lambe, P. C. and Hansen, L. A. (1990). "Design and Performance of Earth Retaining Structures
- Proceedings of a Conference." Geotechnical Special Publication(25 1990): 439-470.
Larson, K. M., Brgmann, R., Bilham, R. and Freymueller, J. T. (1999). "Kinematics of the
India-Eurasia collision zone from GPS measurements." Journal of Geophysical Research
B: Solid Earth 104(B1): 1077-1093.
Lawrence, W. R. (1967). The valley of Kashmir. Kesar Publishers: 478

Lee, I. K., White, W. and Ingles, O. G. (1983). Geotechnical Engineering. Boston, Pitman.
Liao, H. W. (2000). Landslides triggered by Chi-Chi earthquake. National Central University.
Chung-Li: 90.
Lin, J. S. and Whitman, R. V. (1983). "Decoupling approximation to the evaluation of
earthquake-induced plastic slip in earth dams." Earthquake Engineering and Structural
Dynamics 11(5 , Sep.-Oct. 1983, p.667-678.).
Lungarini, L., Troise, C., Meo, M. and De Natale, G. (2005). "Finite element modelling of
topographic effects on elastic ground deformation at Mt. Etna." Journal of Volcanology
and Geothermal Research 144(1-4 SPEC. ISS.): 257-271.
Makdisi, F. I. and Seed, H. B. (1978). "Simplified procedure for estimating dam and
embankment earthquake-induced deformations." Journal of the Geotechnical Engineering
Division 104(GT7): 849-867.
Marcuson, W. F. I. and Curro, J. R. (1981). "Field and laboratory determination of soil moduli."
Journal of the Geotechnical Engineering Division 107(GT10): 1269-1291.
Marinos, V., Marinos, P. and Hoek, E. (2005). "The geological strength index: Applications and
limitations." Bulletin of Engineering Geology and the Environment 64(1): 55-65.
Matasovic, N., Kavazanjian E, Jr. and Liping, Y. (1997). "Newmark deformation analysis with
degrading yield acceleration." Geosynthetics '97. Proc. conference, Long Beach, 1997.
Vol.2: 989-1000.
Mavonga, T. and Durrheim, R. J. (2009). "Probabilistic seismic hazard assessment for the
Democratic Republic of Congo and surrounding areas." South African Journal of
Geology 112(3-4): 329-342.

174

McCrink, T. P. and Real, C. R. (1996). Evaluation of the Newmark method for mapping
earthquake-induced landslide hazards in the Laurel 7.5' Quadrangle, Santa Cruz County,
California. Final Technical Report to the US Geological Survey, Award No. 1434-93-G
2334: 32 pp.
McDougall, J. W. and Khan, S. H. (1990). "Strike-slip faulting in a foreland fold-thrust belt: the
Kalabagh Fault and western Salt Range, Pakistan." Tectonics 9(5): 1061-1075.
McGuire, R. K. (1995). "Probabilistic seismic hazard analysis and design earthquakes: closing
the loop." Bulletin - Seismological Society of America 85(5): 1275-1284.
Meunier, P., Hovius, N. and Haines, A. J. (2007). "Regional patterns of earthquake-triggered
landslides and their relation to ground motion." Geophysical Research Letters 34(20).
Mezcua, J., Rueda, J. and Blanco, R. M. G. (2011). "A new probabilistic seismic hazard study of
Spain." Natural Hazards 59(2): 1087-1108.
Monalisa, Khwaja, A. A. and Jan, M. Q. (2007). "Seismic hazard assessment of the NW
Himalayan fold-and-thrust belt, Pakistan, using probabilistic approach." Journal of
Earthquake Engineering 11(2): 257-301.
Monalisa and Qasim Jan, M. (2010). "Geoseismological study of the Ziarat (Balochistan)
earthquake (doublet?) of 28 October 2008." Current Science 98(1): 50-57.
Mukhopadhyay, D. K. and Mishra, P. (2005). "A balanced cross section across the Himalayan
frontal fold-thrust belt, Subathu area, Himachal Pradesh, India: Thrust sequence,
structural evolution and shortening." Journal of Asian Earth Sciences 25(5): 735-746.
Nason, R. D. (1971). Shattered earth at Wallaby Street, Sylmar, in the San Fernando, California,
earthquake of February 9, 1971. U.S. Geological Survey Professional Paper 733: 97-98.
Newmark, N. M. (1965). "Effects of earthquakes on dams and embankments." Geotechnique
15(2): 139-&.
NORSAR and PMD (Pakistan Meterological Department) (2006). Seismic Hazard Analysis for
the Cities of Islamabad and Rawalpindi.
Nowroozi, A. A. (1985). "Empirical relations between magnitude and fault parameters for
earthquakes in Iran." Bull. Seismol. Soc. Am 75(5): 1327-1338.
Nozu, A., Uwabe, T., Sato, Y. and Shinozawa, T. (1997). Relation between seismic coefficient
and peak ground acceleration estimated from attenuation relation. Technical Note of the
Port and Harbor Research Institute. Ministry of Transport, Japan.
Owen, L. A., Kamp, U., Khattak, G. A., Harp, E. L., Keefer, D. K. and Bauer, M. A. (2008).
"Landslides triggered by the 8 October 2005 Kashmir earthquake." Geomorphology
94(1-2): 1-9.
Petersen, M. D. (2008). Documentation for the 2008 Update of the United States National
Seismic Hazard Maps. USGS.
Petersen, M. D., Dewey, J., Hartzell, S., Mueller, C., Harmsen, S., Frankel, A. D. and Rukstales,
K. (2004). "Probabilistic seismic hazard analysis for Sumatra, Indonesia and across the
Southern Malaysian Peninsula." Tectonophysics 390(1-4): 141-158.
PMD (Pakistan Meterological Department) and NORSAR (2006). Seismic Hazard Analysis and
Zonation of Azad Kashmir and Northern Areas of Pakistan.
Power, M., Chiou, B., Abrahamson, N., Bozorgnia, Y., Shantz, T. and Roblee, C. (2008). "An
overview of the NGA project." Earthquake Spectra 24(1): 3-21.
Pyke, R. (1991) Selection of seismic coefficients for use in pseudo-static slope stability analyses.

175

Quittmeyer, R. C. and Jacob, K. H. (1979). "HISTORICAL AND MODERN SEISMICITY OF
PAKISTAN, AFGHANISTAN, NORTHWESTERN INDIA, AND SOUTHEASTERN
IRAN." Bulletin of the Seismological Society of America 69(3): 773-823.
Raghukanth, S. T. G. (2008). "Ground motion estimation during the Kashmir earthquake of 8th
October 2005." Natural Hazards 46(1): 1-13.
Rathje, E. M. and Bray, J. D. (2000). "Nonlinear coupled seismic sliding analysis of earth
structures." Journal of Geotechnical and Geoenvironmental Engineering 126(11): 1002-
1014.
Reiter, L. (1990). Earthquake hazard analysis : issues and insights. Columbia University Press.
New York: x, 254.
Sanchezsesma, F. J., Herrera, I. and Aviles, J. (1982). "A boundary method for elastsic wave
diffraction- application to scattering of SH-waves by surface irregularities." Bulletin of
the Seismological Society of America 72(2): 473-490.
Sato, H. P., Hasegawa, H., Fujiwara, S., Tobita, M., Koarai, M., Une, H. and Iwahashi, J. (2007).
"Interpretation of landslide distribution triggered by the 2005 Northern Pakistan
earthquake using SPOT 5 imagery." Landslides 4(2): 113-122.
Searle, M. P. (1996). "Geological evidence against large-scale pre-Holocene offsets along the
Karakoram Fault: Implications for the limited extrusion of the Tibetan plateau."
Tectonics 15(1): 171-186.
Seeber, L. and Armbruster, J. (1981). Great detachment earthquakes along the HImalayan Arc
and long-term forecasting. In: Earthquake prediction; an international review. American
Geophysical Union. Washington, DC, United States: pp. 259-277.
Seed, H. B. (1979). "Considerations in the earthquake-resistant design of earth and rockfill
dams." Geotechnique 29(3): 215-263.
Sella, G., Dixon, T. H. and Mao, A. (2002). "REVEL: A model for recent plate velocities from
space geodesy." Journal of Geophysical Research B: Solid Earth 107(4): 11-11.
Singh, D. D. (2000). "Seismotectonics of the Himalaya and its vicinity from centroid-moment
tensor (CMT) solution of earthquakes." Journal of Geodynamics 30(5): 507-537.
Sitar, N. and Clough, G. W. (1983). "Seismic response of steep slopes in cemented soils."
Journal of Geotechnical Engineering-Asce 109(2): 210-227.
Slemmons, D. B. (1982). Determination of design earthquake magnitudes for microzonation.
Proceedings, 3rd Intermational Earthquake Microzonation Conference. Seattle,
Washington. 1: 119-130.
Slemmons, D. B., Bodin, P. and Zhang, X. (1989). "Determination of earthquake size from
surface faulting events." Proc. Int. Seminar on Seismic Zoning, China.
Smith, W. D. (1975). "Application of finite element analysis to body wave progagation
problems." Geophysical Journal of the Royal Astronomical Society 42(2): 747-768.
Somerville, P. G., Smith, N. F., Graves, R. W. and Abrahamson, N. A. (1997). "Modification of
empirical strong ground motion attenuation relations to include the amplitude and
duration effects of rupture directivity." Seismological Research Letters 68(1): 199-222.
Stamatopoulos, C. A., Bassanou, M., Brennan, A. J. and Madabhushi, G. (2007). "Mitigation of
the seismic motion near the edge of cliff-type topographies." Soil Dynamics and
Earthquake Engineering 27(12): 1082-1100.
Stepp, J. C. (1972). Analysis of completeness of the earthquake sample in the Puget Sound Area
and its effect on statistical estimates of earthquake hazard. Proc. of the International
Conference on Microzonations. 2: 897-910.

176

Tang, C. L., Hu, J. C., Lin, M. L., Angelier, J., Lu, C. Y., Chan, Y. C. and Chu, H. T. (2009).
"The Tsaoling landslide triggered by the Chi-Chi earthquake, Taiwan: Insights from a
discrete element simulation." Engineering Geology 106(1-2): 1-19.
Terzaghi, K. (1950). Mechanism of Landslides. The Geological Society of America. Berkey.
November: 83-123.
Tocher, D. (1958). "Earthquake energy and ground breakage." Bulletin of the Seismological
Society of America 48(2): 147-153.
U.S. Department of Energy (U.S. DOE) (1996). Natural Phenomena Hazards Assessment
Criteria, DOE-STD-1023-96. Washington, D.C.
U.S. Nuclear Regulatory Commission (U.S. NRC) (1997). Identification and characterization of
seismic sources and determination of safe shutdown earthquake ground motion,
Regulations 10 CFR Part 100, Regulatory Guide 1.165, Apendix C. Washinton, D.C.
USGS (2010). "Magnitude 7.6 - Pakistan." Earthquake Hazards Program.
http://earthquake.usgs.gov/earthquakes/eqinthenews/2005/usdyae/usdyae.php.
Vilanova, S. P. and Fonseca, J. F. B. D. (2007). "Probabilistic seismic-hazard assessment for
Portugal." Bulletin of the seismological society of America 97(5): 1702-1717.
Wallace, R. E. (1970). "Earthquake recurrence intervals on the San Andreas fault." Geological
Society of America Bulletin 81: 2875-2890.
Weichert, D. H. (1980). "Estimation of the earthquake recurrence parameters for unequal
observation periods for different magnitudes." Bulletin of the seismological society of
America 70: 1337-1356.
Wells, D. L. and Coppersmith, K. J. (1994). "New empirical relationships among magnitude,
rupture length, rupture width, rupture area, and surface displacement." Bulletin of teh
Seismological Society of America 84(4): 974-1002.
WGCEP (2003). Earthquake probabilities in the San Francisco Bay Region 2002-2031.
WGCEP (2008). The uniform California earthquake rupture forcast, version 2 (UCERF 2). U.S.
Geological Survey Open-File Report 2007-1437 and California Geological Survey
Special Report 203.
Yegian, M. K., Marciano, E. A. and Ghahraman, V. G. (1991). "Earthquake-induced permanent
deformations. Probabilistic approach." Journal of geotechnical engineering 117(1): 35-50.
Yin, J., Chen, J., Xu, X., Wang, X. and Zheng, Y. (2010). "The characteristics of the landslides
triggered by the Wenchuan Ms 8.0 earthquake from Anxian to Beichuan." Journal of
Asian Earth Sciences 37(5-6): 452-459.
Yin, Y., Wang, F. and Sun, P. (2009). "Landslide hazards triggered by the 2008 Wenchuan
earthquake, Sichuan, China." Landslides 6(2): 139-151.
Youngs, R. R. and Coppersmith, K. J. (1985). "Implications of fault slip rates and earthquake
recurrence models to probabilistic seismic hazard assessments." Bulletin of the
seismological society of America 75(4): 939-964.
Zhao, C. and Valliappan, S. (1993). "Seismic wave scattering effects under different canyon
topographic and geological conditions." Soil Dynamics and Earthquake Engineering
12(3): 129-143.

You might also like