historical and modern seismotectonics of the indian plate with

193
Historical and Modern Seismotectonics of the Indian Plate with an Emphasis on its Western Boundary with the Eurasian Plate by W. M. Szeliga B. S., University of Massachusetts, 2003 M. S., Central Washington University, 2005 A thesis submitted to the Faculty of the Graduate School of the University of Colorado in partial fulfillment of the requirements for the degree of Doctor of Philosophy Department of Geological Sciences 2010

Upload: khangminh22

Post on 19-Jan-2023

0 views

Category:

Documents


0 download

TRANSCRIPT

Historical and Modern Seismotectonics of the Indian Plate with anEmphasis on its Western Boundary with the Eurasian Plate

by

W. M. Szeliga

B. S., University of Massachusetts, 2003

M. S., Central Washington University, 2005

A thesis submitted to the

Faculty of the Graduate School of the

University of Colorado in partial fulfillment

of the requirements for the degree of

Doctor of Philosophy

Department of Geological Sciences

2010

This thesis entitled:Historical and Modern Seismotectonics of the Indian Plate with an Emphasis on its Western Boundary

with the Eurasian Platewritten by W. M. Szeliga

has been approved for the Department of Geological Sciences

Roger Bilham

Peter Molnar

Date

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

iii

Szeliga, W. M. (Ph. D., Geophysics)

Historical andModern Seismotectonics of the Indian Platewith an Emphasis on itsWestern Boundarywith

the Eurasian Plate

Thesis directed by Dr Roger Bilham

The western edge of the Indian plate is a transform plate boundary similar to the San Andreas

Fault in that it lies mostly on land, has a similar expected slip rate, accommodates restraining bends, and

contains segments that may slip aseismically by surface creep. Tectonic models of the western edge of

India must also account for the absence of significant seismic moment release in the past century along

the Chaman Fault, the transform boundary between Asia and India. I discuss modern and historical data

from India and Pakistan that provide new constraints on deformation within this 100–250 km wide plate

boundary. Geological andplate-closure estimates suggest sinistral slip of 19–35mm/yr since theOligocene

across the Chaman Fault system. Analysis of space-based geodetic data suggests a prevalence of shallow

locking depths and an upper limit of approximately 19.5 mm/yr of sinistral motion across the Chaman

Fault System south of Afghanistan. In the past century, the region between the Chaman Fault System and

the Indus Plain near Quetta, Pakistan, has experienced numerous earthquakes with a larger total moment

release than an equivalent length of the Himalaya in the same period, comparable to a single Mw 8.0.

Of this moment release, 90% has occurred more than 70 km east of the Chaman fault. In this region, GPS

data have captured slip partitioning across the plate boundary suggesting that long-term sinistral slip is

shared between the Chaman and Ghazaband fault systems. Additionally, a combination of GPS and InSAR

analysis of a pair of Mw 6.4 earthquakes NE of Quetta in 2008 suggests that they occurred on a parallel

pair of sinistral faults, rather than the dextral mechanism suggested by their NW-SE trending fault planes.

I find that “bookshelf faulting” occurs in a zone NE of Quetta that includes several previous instrumental

and historical earthquakes. This geodetic view of deformation in Pakistan differs from that derived from

the instrumental seismic record, but is consistent with the sparse historical record of earthquakes in the

past two millennia, and has important implications for assessment of seismic hazards in Pakistan.

Dedication

To My Family

v

Acknowledgements

The compilation of macroseismic intensity data presented in Chapter 2 was in part supported by a

library research grant from Munich Re and conducted by Stacey Martin. Analysis of macroseismic inten-

sity data performed in Chapters 2 and 3 was funded through National Science Foundation grant number

EAR-00004349. Material in Chapter 4 is based on research supported by the National Science Foundation

under grant number EAR-0229690. Research presented in Chapter 5 was funded by the National Science

Foundation under grants EAR-003449 and EAR-0739081. Research presented in Chapter 6 was supported

by the National Science Foundation under grant number EAR-0729081.

ERS and Envisat data were provided by the European Space Agency under a category-1 proposal

number 2757 and were processed using the JPL/Caltech software package ROI PAC. Original InSAR data

are copyright of the European Space Agency.

Occupation andmaintenance of continuous and campaignGlobal Positioning System (GPS) receivers

was performed byDinMohammadKakar of theUniversity of Baluchistan, Quetta, Pakistan and Sarosh Lodi

of the NED University, Karachi, Pakistan. GPS data were processed using the MIT GAMIT/GLOBK software

package.

Seismic waveform data presented in Chapter 6 was obtained from Global Seismic Network sta-

tions using Iris’s Wilbur II system (www.iris.edu/wilbur). Data from the Global Centroid Moment Tensor

Project (CMT) was retrieved from http://www.globalcmt.org. Monthly Hypocenter Data File (MHDF)

data were retrieved using SeismiQuery (www.iris.edu/dms/sq.htm). International Seismological Centre

(ISC) and Engdahl, van der Hilst and Buland (EHB) catalog data were retrieved from http://www.isc.

ac.uk. In addition, I would like to acknowledge the following seismic networks: the Alaska Regional Net-

vi

work, the Australian Seismological Centre, the CanadianNational SeismicNetwork, the Czech SeismicNet-

work, GEOSCOPE, GEOFON, the Global Telemetered Southern Hemisphere Network, the IRIS China Digital

Seismic Network, the IRIS/IDA Network, the IRIS/USGS Network, the Japan Meteorological Agency Seis-

mic Network, MEDNET, the Malaysian National Seismic Network, the Polish National Seismic Network,

the Portuguese National Seismic Network, and the Broadband Array in Taiwan for Seismology. Landsat

imagery was acquired using the US Geological Survey’s Earth Explorer (http://edcsns17.cr.usgs.gov/

EarthExplorer).

Figures 5.3 and 5.4 were provided by Dr. Daniel Schelling of Structural Geology LLC, Salt Lake City,

UT. Figure 3.9 was created using gnuplot (http://www.gnuplot.info), all other figures were created using

Generic Mapping Tools (Wessel and Smith, 1998). I would like to thank Dr. Robert McCaffery of Rensselaer

Polytechnic Institute for discussion on teleseismic data processing, Drs. Eric Fielding of the Jet Proplusion

Laboratory, Gareth Funning of the University of California, Riverside, Rowena Lohman of Cornell Univer-

sity and Tim Wright of the University of Leeds for discussion on InSAR processing techniques. I would

also like to thank Dr. Gareth Funning for providing software for InSAR inversion and Dr. Rowena Lohman

for providing software for InSAR data resampling. I would like to thank Dr. Daniel Schelling for providing

detailed geological and structural information about Baluchistan. For numerous interesting discussions

and reviews I am indebted to Karl Mueller and Andrew Meigs. Miriam Garcia’s SOARS/RECESS internship

during the Summer of 2006 was instrumental in preliminary modeling that led to the results discussed in

Chapter 5.

vii

Contents

Chapter

1 Introduction 1

1.1 The Indian Plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Geodetic and Seismic Observations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.3 Thesis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

2 A Catalog of Felt Intensity Data for 570 Earthquakes in India from 1636 to 2009 7

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2.2 Intensity Scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2.3 Reporting Consistency and Completeness . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.4 Summary of Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

2.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

3 Intensity, Magnitude, Location and Attenuation in India for Felt Earthquakes since 1762 21

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

3.2 Data and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

3.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

3.3.1 Comparisons with previous attenuation studies . . . . . . . . . . . . . . . . . . . 27

3.4 Estimation of Historical Epicenters and Magnitudes . . . . . . . . . . . . . . . . . . . . . 35

3.4.1 Epicentral Locations and Magnitudes of Historical Events . . . . . . . . . . . . . . 35

viii

3.4.2 Catalog completeness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

3.5 Case Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

3.5.1 The 1803 Uttarakhand Himalaya Earthquake . . . . . . . . . . . . . . . . . . . . . 39

3.5.2 The 1819 Allah Bund Earthquake . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

3.5.3 The 1833 and 1866 Nepal Earthquakes . . . . . . . . . . . . . . . . . . . . . . . . . 45

3.5.4 The 2001 Bhuj Earthquake (Mw 7.6) . . . . . . . . . . . . . . . . . . . . . . . . . . 47

3.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

3.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

4 Interseismic Strain Accumulation along the Western Boundary of the Indian Subcontinent 54

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

4.2 Tectonic Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

4.3 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

4.3.1 GPS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

4.3.2 InSAR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

4.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

4.4.1 Ornach-Nal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

4.4.2 Chaman Fault near Chaman . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

4.4.3 Chaman Fault near Qalat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

4.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

4.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

5 Fold and thrust partitioning in a contracting fold belt: Insights from the 1931 Mach Earthquake in

Baluchistan 77

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

5.2 Structural setting of the Bolan Pass Region . . . . . . . . . . . . . . . . . . . . . . . . . . 80

5.3 GPS measurements of convergence and shear between the Asian and Indian Plates . . . . 83

5.4 Macroseismic location of the Mach earthquake . . . . . . . . . . . . . . . . . . . . . . . . 85

ix

5.5 Leveling data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

5.6 Discussion: the earthquake cycle in a ramp-flat-ramp system . . . . . . . . . . . . . . . . 90

5.7 Geodetic convergence, slip potential and renewal time . . . . . . . . . . . . . . . . . . . . 94

5.8 Sequential triggering of ruptures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

5.9 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

6 Bookshelf Faulting in the 2008 Ziarat Earthquake Sequence, Northern Baluchistan 101

6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

6.2 Tectonic Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

6.3 Data and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

6.3.1 Double-difference Relocations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

6.3.2 Teleseismic Body-wave Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

6.3.3 InSAR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

6.3.4 GPS Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

6.3.5 Macroseismic Observations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

6.4 Interpretational Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

6.4.1 The 9 Dec. 2008 Aftershock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

6.4.2 28–29 Oct. Mainshocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

6.4.3 16 Nov. 1993 Earthquake . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

6.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

6.5.1 Historical Seismicity and Shear Zone Extent . . . . . . . . . . . . . . . . . . . . . 125

6.5.2 Shear Zone Seismic Productivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

6.5.3 Tectonic Analogues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

6.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131

7 Conclusions 133

7.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133

7.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138

x

Appendix

A EMS-98 Short Form 140

B List of Epicentral Locations for Historical Seismicity on the Indian Plate 142

Bibliography 151

xi

Tables

Table

2.1 The first dozen earthquakes from the electronic supplement toMartin and Szeliga (2010) to

illustrate format. Columns Year, Month and Day refer to the date of an event in local time.

For earthquakes with more than seven intensity observations (column Number of Obser-

vations), the approximate epicentral location is listed (columns Longitude, Latitude). The

number of observations corresponds to the number of intensity reports listed in the elec-

tronic supplement to Martin and Szeliga (2010). A geographic region designator is defined

for some events (column Earthquake). This column serves as a reference column to groups

of intensity observations in Table 2.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.2 The first 5 earthquakes of 570 from the electronic supplement toMartin and Szeliga (2010).

Columns Year, Month and Day refer to the date of an event in local time. Columns Longi-

tude and Latitude refer to the location of the intensity observation. Column EMS-98 lists

assessed EMS-98 intensities (Grunthal and Levret, 2001). The geographic location of each ob-

servation is listed in column Location. Column Earthquake serves to group observations

from the same earthquake and refers to the geographic location of each earthquake in Ta-

ble 2.1. Earthquakes with fewer than 2 observations are not assigned geographic locations. 9

2.3 Regression coefficients and anticipated mean return time in years for shaking at EMS-98

intensities V, VI and VII for the five largest cities in India. . . . . . . . . . . . . . . . . . . 17

3.1 Intensity attenuation relationship parameters for India, the Indian Craton and the Hi-

malaya. Columns a, b, c, and d refer to the variables in equation (3.1). . . . . . . . . . . . . 26

xii

3.2 Intensity attenuation relationship coefficients obtained by other investigations used in

this paper. Columns a, b, c, and d refer to the variables in equation (3.1). The form of the

attenuation relationship used by Atkinson and Wald (2007) and its associated coefficients

are listed in Table (1) and equation (1) in Atkinson and Wald (2007). (a) This parameter was

defined to be zero. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

3.3 Epicentral locations and intensity magnitudes (MI ) of the 1819 Allahbund earthquake de-

termined using the method outlined in Section 3.2. Uncertainty in descriptions of damage

to the towns of Baliari and Umarkot in MacMurdo (1823) permit a range of EMS-98 inten-

sities with a resulting range in the epicentral location and magnitude for the 1819 earth-

quake (Figure 3.11). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

5.1 Instrumental and inferred macroseismic locations for the three earthquakes. . . . . . . . 85

5.2 Observed (“Obs”) and synthetic slip on the decollement. Segments are free to slip in re-

sponse to 10mof thrust displacement imposed on the deepest fault segment, a value scaled

to approximate themean observed coseismic slip. “Co-8” refers to coseismic slip shallower

than approximately 8 km depth, and “Co-9” refers to coseismic slip from one segment

deeper at approximately 9 km depth. “No-Lock” indicates the slip that would occur in the

absence of interseismic locking, and “interseismic” indicates the synthetic slip that occurs

below a locking line at 9 km depth. I have scaled the driving element to 10 km so that that

synthetic slip approximates the mean slip derived from the observed leveling data. . . . . 91

xiii

5.3 Calculations of partitioned convergence. Geometric relations between applied geodetic

displacement and slip on theDezghat/Bannh thrust fault for a range of hypothetical decoll-

ment depths (the actual depth is believed to lie in the range 18–20 km). The imposed dis-

placement, S, is that calculated to cause the mean observed coseismic slip, s, in the Mach

earthquake. D is the mean depth of the decollement, and d, is the approximate starting

depth of the frontal thrust above a steeper ramp connecting the two. The ratio S/s is a

proxy for the increase in the recurrence interval for earthquakes on the frontal thrusts

compared to the time that would be calculated from geodetic convergence rates of the

entire range. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

6.1 Historical earthquakes in the Quetta Syntaxis. An additional 5 earthquakes with magni-

tudes between Mw 5.1 and Mw 5.4 occurred during the Oct.–Dec. 2008 aftershock se-

quence but are unlisted. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

6.2 Comparison of the fault plane parameters for the preferred double-couple rupture plane of

the 9Dec. 2008 aftershock. Strike, dip, rake anddepth are for the preferred fault plane from

the double-couple with the largest contribution to the total moment. Moment is the total

moment of the entire event. Each inversionmethod is sensitive to deformation in different

frequency bands. To illustrate this, solutions are arranged vertically from shortest (Body-

wave) to longest (InSAR) period of sensitivity to radiated energy. For a visual comparison

of each solution, see Figure 6.7. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

6.3 Comparison of fault plane parameters for the preferred double couple rupture plane of the

28 Oct. 2008 mainshock. Strike, dip, rake and depth are for the fault plane with the largest

moment release. Moment is the total moment of the entire event. Solutions are arranged

vertically from shortest (Body-wave) to longest (InSAR) period. For a visual comparison

of each solution, see Figure 6.7. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

xiv

6.4 Comparison of fault plane parameters for the preferred rupture plane of the 29 Oct. 2008

mainshock. Strike, dip, rake and depth are for the fault plane with the largest moment re-

lease. Moment is the total moment of the entire event. Solutions are arranged vertically

from shortest (Body-wave) to longest (InSAR) period. For a visual comparison of each so-

lution, see Figure 6.7. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120

6.5 Comparison of fault plane parameters for the preferred rupture plane of the 16 Nov. 1993

earthquake. Strike, dip, rake and depth are for the fault plane with the largest moment

release. Moment is the total moment of the entire event. Each inversion method is sensi-

tive to deformation in different frequency bands. To illustrate this, solutions are arranged

vertically from shortest (CMT) to longest (InSAR) period of sensitivity to radiated energy.

For a visual comparison of each solution, see Figure 6.7. . . . . . . . . . . . . . . . . . . . 123

A.1 The short form of the EMS-98 intensity scale reproduced from Grunthal and Levret (2001).

For a more detailed description of the criteria used to assign intensities, refer to Grunthal

and Levret (2001), specifically pages 14–20. . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

xv

Figures

Figure

2.1 A cumulative histogram of earthquakes per 50 year period in the historical seismic catalog

(right hand axis). Vertical bars topped with circles (left hand axis) show observations per

earthquake. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.2 (a) Circles indicate the locations of intensity data listed in the electonic supplement toMar-

tin and Szeliga (2010). Regions with low population density, such as the Rajasthan desert,

parts of Baluchistan, theNepal andAssamHimalaya and the Indo-Burman ranges are poorly

represented historically. Communication routes and rail lines show up as faint lines in the

data. (b) Epicenters for historic earthquakes listed in the electronic supplement to Martin

and Szeliga (2010) determined using the method of Bakun and Wentworth (1997). . . . . . . 14

2.3 (a). Maximum shaking intensity observed during the period 1636–2009. (b). Interpolated

maximum shaking intensity observed during the period 1636–2009. (c). Interpolatedmax-

imum shaking intensity in Gujarat. (d). Map of average shear wave velocity down to 30

m (Vs30) for the Indian state of Gujarat. (e). Interpolated maximum shaking intensity

in northeast India. (f). Vs30 map of the northeastern India. In producing interpolated

maximum shaking intensity maps, locations within 10 km of one another were binned to

account for differences in location names and centers of population over time. Maximum

shaking intensity data were interpolated using a nearest neighbor schema. Vs30 maps

were derived from 30 arcsecond SRTM V 2.0 data (Farr et al., 2007) using the techniques

outlined inWald and Allen (2007). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

xvi

2.4 (a) Cumulative number of earthquakes felt inmajor Indian cities since 1762. (b) Frequency

of maximum shaking intensities observed in these cities in the past two hundred years.

The regression coefficients to these data, fit between intensity II and V are shown in Table

2.3. The well behaved form of these curves suggests that the probability for future shaking

frommodest earthquakes can be estimated with reasonable confidence. The estimation of

the probable return time of higher intensity shaking from these curves is less well con-

strained. The light gray line is the regression line for Delhi using the coefficients from

Table 2.3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

3.1 Epicentral locations of 29 calibration events. I have excluded earthquakes with depths in

excess of 40 km. Events marked with diamonds were used to determine cratonic attenua-

tion while events marked with circles were used to determine Himalayan attenuation. . . 25

3.2 Intensity distributions for the data used to calculate the attenuation parameters in Table

(3.1). (a) Distance to earthquake centroid versus moment magnitude for events in the Hi-

malaya. (b) Distance to earthquake centroid versus moment magnitude for events on the

Craton. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

3.3 Intensity attenuation with distance for a hypothetical M 6.5 Himalayan earthquake from

this study (solid line) and from Ambraseys and Douglas (2004) (dashed line). Intensity data

from this study are in EMS-98 and data from Ambraseys and Douglas (2004) are inMSK. Error

bars are 2σ. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

xvii

3.4 Comparison of assessed intensities at 95 common locations from the catalog and Ambraseys

andDouglas (2004) for 3 earthquakes. For the histogram, thex-axis (top) corresponds to the

normalized frequency of the combined intensity differences. For individual earthquakes,

x-axis (bottom) corresponds to the assessed intensity value from the catalog. The y-axis

corresponds to the difference between the assessed intensities from the catalog and those

from Ambraseys and Douglas (2004) with negative values indicating that the intensity from

the catalog is lower than that listed in Ambraseys and Douglas (2004). For clarity, intensities

for the 1819 Allah Bund and 1833 Nepal earthquake have been artificially offset to the right

by 0.1 and 0.2 intensity units respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

3.5 Intensity attenuation relationship between India from this study, the results of Bakun et al.

(2003) for easternNorth America, and the results ofAtkinson andWald (2007) for the Central

EasternUS (CEUS) for a hypotheticalM 6.5 earthquake. Indian intensity data are in EMS-98

while data from eastern North America are in MMI. Error bars are 2σ. . . . . . . . . . . . 32

3.6 A direct comparison between intensity observations from eastern North American and

cratonic India. Eastern North American intensity data are from the USGS Community In-

ternet Intensity Map Project, error bars represent standard error estimates of the sam-

ple median. a.) Direct comparison of the median distance to which each intensity was

observed for the 18 April 2008 Mw 5.2 Mt. Carmel, IL earthquake and the 5 September

2000 Mw 5.2 Koyna earthquake. For intensities III–VI, the median distance is statistically

larger for the Mt. Carmel, IL earthquake. b.) Direct comparison of the median distance to

which each intensity was observed for the 29 April 2003Mw 4.6 Fort Payne, AL and the 26

November 2007 Mw 4.7 Delhi earthquake. Although the Delhi earthquake is larger than

the Fort Payne earthquake, the median distance to which intensities II–V are smaller in

India. This suggests that the attenuation difference between eastern North American and

India is equivalent to a magnitude increase of at least 0.2Mw. . . . . . . . . . . . . . . . . 33

xviii

3.7 Intensity attenuation relationship between the Himalaya from this study, the results of

Bakun andWentworth (1997) for California, and the results from Atkinson andWald (2007) for

California for a hypothetical M 6.5 earthquake. Indian intensity data are in EMS-98 while

data from California are in MMI. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

3.8 Comparison of the epicentral misfit for instrumentally recorded earthquakes in the Koyna

region of India. On both figures, the arrowpoints from the instrumental epicenter towards

the intensity derived epicenter. (a) Epicentralmisfit in the Koyna region using the location

of the minimum of equation (3.2) as the epicentral estimate. (b) Epicentral misfit in the

Koyna region using the location of the minimum M from equation (3.2). . . . . . . . . . . 37

3.9 Frequency-magnitude plot of earthquakes occurring on the Indian subcontinent. Filled

circles represent events from the ISC catalog during 1980–2000. Diamonds represent events

from the catalog; open circleswith synthetic aftershock sequences added. Dashed line rep-

resents a frequency-magnitude relationship with a b value of 1.0 . . . . . . . . . . . . . . 40

3.10 The location of the 1803 Uttarkashi earthquake as determined by the method outlined in

Section 3.2. The contours represent the 50% and 67% confidence contours as determined

by Bakun (1999). The instrumental epicenters of the 1991 Uttarkashi and 1999 Chamoli

earthquakes (stars) are shown for reference. The location of the 1803 Uttarkashi earth-

quake as determined by Ambraseys and Douglas (2004) is illustrated by a square. I reject

the alternative epicentral location permitted by the data near the Ganges (indicated by

the closed 50% and 67% confidence contours). Filled circles indicate the locations of felt

reports for the 1803 earthquake within 250 km of the epicenter. . . . . . . . . . . . . . . . 42

xix

3.11 Possible locations for the 1819 Allahbund earthquake as determined by the method out-

lined in Section 3.2 (open arrows with calculated Mw). The parameters of these possible

locations are listed in Table 3.3. The location of the fault responsible for the 2001 Bhuj

Mw 7.6 earthquake (Schmidt and Burgmann, 2006) as well as the location of the Allah Bund

fault (Malik et al., 2001) are shown with barbs on the hanging wall. The location of the in-

ferred Island Belt Fault is shown with a dashed line (Malik et al., 2001). Contours represent

magnitudes from the epicentral location algorithm (Section 3.2) using the raw intensity

data; they indicate a minimum magnitude location in the Gulf of Kachchh. The locations

of Umarkot and Baliari are shown for reference. Filled circles represent felt intensity loca-

tions within 300 km of the epicenter and arrows indicate the change in epicentral location

due to changes outlined in Table 3.3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

3.12 The locations of the 1833 and 1866Nepal earthquakes as determined using themethod out-

lined in Section 3.2. The contours represent the 50% and 67% confidence regions obtained

using method described by Bakun (1999). The previous estimate of epicentral location for

the 1833 earthquake from Ambraseys and Douglas (2004) is represented by a square. Filled

circles indicate the locations of felt reports for the 1833 and 1866 earthquakes within 250

km of Kathmandu. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

3.13 Intensity observations of the 2001 Bhuj Mw 7.6 earthquake compared to the attenuation

curve derived for cratonic India for an earthquake of Mw 7.6. Open circles represent ob-

served intensities, diamonds represent the median distance for each observed intensity

level. Dashed lines represent the 2-σ envelope of uncertainty in the intensity attenuation

model as a function of distance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

xx

4.1 Map of the western boundary of the Indian Plate, highlighting the major faults of the

Chaman Fault System, place names mentioned in the text are also indicated. The map

projection is oblique Mercator about the pole of relative motion between the Indian and

Eurasian plates. Thrust faults are shownwith filled triangles on the hanging wall, all other

faults shown are strike slip. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

4.2 Map showing GPS station locations and names along the western boundary of the Indian

Plate. A.) GPS stations throughout Pakistan. Dashed rectangle indicates the ground foot-

print of Envisat track 213 frame 621. B.) Stations in the Quetta Syntaxis where there is a

high station density. C.) Stations along the Makran Coast. . . . . . . . . . . . . . . . . . . 59

4.3 Date versus perpendicular baseline plot for Envisat track 213, frame 621. Filled circles rep-

resent individual SAR scenes and solid lines represent interferograms. There is one per-

pendicular baseline outlier indicated on the correct date in parenthesis along side the as-

sociated perpendicular baseline value. The vertical dashed line corresponds to anMw 5.0

earthquake on 21Oct. 2005 along the Chaman fault in the northern portion of Envisat track

213 frame 621. The 12 interferograms shown have a median perpendicular baseline of 30

m, corresponding to an altitude of ambiguity of more than 450 m. . . . . . . . . . . . . . . 61

4.4 GPS velocities of stations from the Makran region of Pakistan. All velocities are relative to

the stable Indian Plate as defined in Altamimi et al. (2007) and are plotted using a Mercator

projection. The exact location of the offshore intersection of the subduction zone and the

Chaman Fault System is unknown and is denoted with a question mark. Station names

appear on Figure 4.2C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

xxi

4.5 GPS profile across the Ornach-Nal Fault. Velocities and uncertainties are projected into a

direction parallel to the Ornach-Nal Fault and are relative to the stable Indian Plate. Un-

certainties shown are 2σ. The thick horizontal bar indicates the 95% HPD range for pos-

sible fault locations. The dotted line represents the model that maximizes the empirical

posterior likelihood function as determined using a Markov-Chain Monte Carlo method

(Mosegaard and Tarantola, 1995). The slip rate and locking depth for the fault location that

satisfies both the posterior likelihood and geological critera (the nominal plate boundary)

are indicated on the figure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

4.6 Landsat 7 image of the southern Ornach Nal fault and adjacent Hinglaj synform. The large

square brackets indicate the spatial region encompassed by the 95% HPD region shown in

Figure 4.5. Geologically likely locations for the plate bounding fault(s) are indicated by the

NE-SW trending dashed lines. The preferred plate bounding fault is the easternmost left

stepping pair of faults across the Hinglaj synform. The gap between the fault tips corre-

sponds to the deepest part of the synform (Bannert et al., 1992) and is likely a pull apart

feature. GPS velocities are relative to the stable Indian Plate and are identical to those

shown in Figure 4.4. The image is a combination of bands 7, 4 and 2 to highlight geological

information. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

4.7 GPS velocities of stations in the region of Quetta, Pakistan. All velocities are relative to

the stable Indian Plate as defined in Altamimi et al. (2007) and are plotted using a Mercator

projection. Station names appear on Figure 4.2B. . . . . . . . . . . . . . . . . . . . . . . . 66

4.8 GPS profile across the Chaman Fault. Velocities and uncertainties are projected into a

direction parallel to the Chaman Fault and are relative to the stable Indian Plate. Uncer-

tainties shown are 2σ. The dotted line represents the model that maximizes the empirical

posterior likelihood function as determined using a Markov-Chain Monte Carlo method

(Mosegaard and Tarantola, 1995). The slip rate and locking depth for this model are indi-

cated on the figure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

xxii

4.9 InSAR rate-map derived from stacking 12 ascending pass Envisat interferograms. Solid

arrow indicates the flight direction of the satellite while the transparent arrow indicated

the line-of-sight direction. Values are phase velocity in rad/yr in the line-of-sight of the

radar and referenced to a pixel in the far NW corner of the scene. More positive values

of phase velocity indicate increasing radar line-of-sight distance. Interferograms used in

construction of the rate-map are indicated by solid lines in Figure 4.3 and have a median

perpendicular baseline of 30 m. The surface trace of the Chaman Fault is indicated by the

dashed line. For reference, the locations of GPS stations CHMC and SHBG are indicated

in the southern portion of the map. The increasing radar line-of-sight velocities near the

town of Qalat, Afghanistan, (black triangle) are likely tied to subsidence due to groundwa-

ter withdrawal for agriculture. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

4.10 Chaman fault centered profile of line-of-sight velocities from the InSAR rate-map shown in

Figure 4.9. Increasing line-of-sight velocities represent motion away from the radar. The

gray data are SRTM level 2 3s topography sampled in the same manner as the InSAR data.

Larger variances in the topographic data indicate larger changes in topography parallel

to the Chaman fault. The dashed line corresponds to the Monte Carlo derived model. Slip

rate and locking depth are calculated in the radar line-of-sight. The convex-up feature 25

km northwest of the Chaman Fault corresponds with groundwater withdrawal near the

town of Qalat, Afghanistan. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

4.11 Comparison of InSAR short-baseline results and Landsat 7 imagery from the Tarnak Rud

valley near the town of Qalat, Afghanistan. A.) Line-of-sight (LOS) rate map of ground

subsidence near the town of Qalat, Afghanistan (triangle). Positive values indicate motion

away from the radar. Solid arrow indicates the flight direction of the satellite and outlined

arrow denotes the line-of-sight direction of the satellite. Black trianglemarks the location

of the town of Qalat and is the same as in Figure 4.9. B.) Landsat 7 image from 18 May 2003

using band combination 4,3,2 to highlight vegetation (red areas). . . . . . . . . . . . . . . 71

xxiii

4.12 Maximum fault-normal and fault-parallel velocities based on ITRF05 pole-of-rotation lo-

cations and rates published in Altamimi et al. (2007) projected along mapped faults on the

western boundary of the Indian Plate. Estimates are derived using the azimuth of the sur-

face trace of plate bounding faults and assume that slip partitioning is perfect and occurs

only along a single fault. A.) Maximum convergence estimated assuming perfect partition-

ing of slip. Locations of convergence observations indicated by text. B.)Maximum sinistral

motion estimated assuming perfect partitioning of slip. Locations of sinistral motion esti-

mates indicated by horizontal bars and represent 95% confidence intervals. Note all three

measurements of sinistral motion and bothmeasurements of fault-normalmotion suggest

lower rates compared with perfect slip partitioning. . . . . . . . . . . . . . . . . . . . . . 75

5.1 A.) Recent seismicity (Mw > 5) and instrumental locations for the Sharigh, Mach (stars)

andQuetta earthquakes (focalmechanismbeachball) and their inferred causal faults (Quetta

rupture dashed and Bannh fault shown as surface thrust NE of the instrumental epicenter).

Focal mechanisms scaled according to magnitude - the largest focal mechanism isMw 7.7

(Singh and Gupta, 1980) and the smallest isMw 5 (all from the Global CMT). B.) Interpolated

Intensity VIII isoseismals for the three earthquakes, the path of the 1909–1936 leveling

line and GPS velocity vectors 2005–8 relative to fixed India. The approximate rupture zone

of the Mach earthquake is shown by the rectangle. The intensity-derived epicenters are

shown on each map as a star. The Quetta centroid solution lies at the opposite end of the

rupture from the intensity solution as a result of directivity. . . . . . . . . . . . . . . . . . 79

xxiv

5.2 Schematic sections of vertical deformation and subsurface geometry of previous attempts

to emulate observed uplift data in the Mach earthquake (Figures 5.2(a) and 5.2(b)). These

models invoked uniform subsurface slip on shallow east-dipping planar thrusts. In Figure

5.2(a) planar, uniform slip is invoked with no structural control (Ambraseys and Bilham,

2003a). In Figure 5.2(b) the speculative wedge thrust geometry of Banks and Warburton

(1986) constrains two fault planes on which combinations of uniform slip were imposed

to obtain the best-fitting surface uplift (Garcia et al., 2006). Spatially variable slip on the

west-dipping Bannh fault (Bannert et al., 1992; Schelling, 1999a) is presented here (2c). . . . 79

5.3 Geological map of the Bolan Pass region of the northern Kirthar Range, showing the loca-

tions of the balanced structural cross section and leveling line discussed in the paper. . . 81

5.4 Balanced structural cross section across the deformation front of the northern Kirthar

Range in the Bolan Pass area and in the vicinity of the leveling line. See Figure 3 for cross

section location and text for discussion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

5.5 GPS velocities projected E-W showing westward velocities relative to stable India. For lo-

cations, see Figure 5.1B. The GPS points, with one exception, show convergence with fixed

India at 5± 1mm/yr. The one exception is QTAG, the continuous GPS point at Quetta. . . 84

5.6 Macroseismc epicenters for the Sharigh, Mach and Quetta earthquakes. The dashed con-

tours in this figure are not isoseismals but iso-magnitude contours using the method of

Bakun and Wentworth (1997). They indicate the required magnitude for each earthquake

had it been located on these contours. The preferredmacroseismic epicentral location lies

within the closed contour of theminimum-variance solution shown as solid lineswhile the

stars represent the instrumentally located epicenters. . . . . . . . . . . . . . . . . . . . . 86

5.7 Leveling data, topographic relief and subsurface section simplified from Figure 5.4. The

synthetic fit to the data results from spatially varying slip on the Dezghat and Bannh faults

(dashed line on section). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

xxv

5.8 Geometry of the active decollement and frontal thrust (bold line with depth scale right),

and inferred slip on segments shallower than 9 km (grey envelope) compared to synthetic

slip (slip scale left). The calculated slip for the entire fault is given by the top staircase-

line (21, 3-km-long freely-slipping segments responding to an input displacement of 10 m

imposed from the left (west)). The lower staircase-lines are formed from two calculations:

slip anticipated below a locking line at 9 km depth (synthetic pre-seismic slip), and the slip

during rupture at shallower depths that occurs when this interseismic slip distribution

drives co-seismic rupture (synthetic co-seismic). The difference between the two lower

staircase lines and the upper staircase is the slip deficit caused by interseismic locking at

9 km depth. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

5.9 Space-time history of seismic moment release as a function of distance from the inferred

Mach 1931 earthquake rupture zone. More than 89%of the total seismicmoment release in

the past 200 years (within a radius of 500 km centered on the Mach earthquake) occurred

between 1931 and 1935. All known earthquakes larger than M6.5 are included in this plot. 97

6.1 Map of the Sulaiman Lobe and northern Kirthar Range of Pakistan, highlighting the major

faults of region. The Bannh and Dezghat faults last ruptured during the 1931 Mach earth-

quake. The Ghazaband Fault is presumed to have last ruptured during the 1935 Quetta

Earthquake and the Chaman Fault last ruptured in 1892 and 1976. The Katawaz Block of

Haq and Davis (1997) is outlined with a dashed line. The three stars indicate the locations

of the twomainshocks and the largest aftershock of the 2008 Pishin Earthquake sequence.

The Kingri Fault is a sinistral fault and is presumed to enable the southward extrusion of

the Sulaiman Lobe (Rowlands, 1978). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

6.2 Map of the Sulaiman Lobe and northern Kirthar Range of Paksitan, showing the loca-

tion of towns mentioned in the text. Historical earthquakes in the Quetta Seismic zone,

1900–2010. Numbers reference dates, epicenters and magnitude listed in Table 6.1. Only

the three largest earthquakes from the 2008 aftershock sequence are shown on the map. . 104

xxvi

6.3 Spatially averaged GPS velocities with respect to the stable Indian Plate and centroid mo-

ment tensors from the Global CMT (Dziewonski et al., 1981) withMw > 5 since 1976. Filled

regions are compressional quadrants of the best-fitting double couple. Note the lack of

seismicitywithin the boundaries of the Katawaz Block (see Figure 6.1 for place names). Ve-

locities are calculated as the weighted spatial average of all regional GPS velocities within

a 30’ grid. The location of each velocity average is calculated as the mean of the locations

within each grid. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

6.4 Scene acquisition date versus perpendicular baseline for Envisat track 213 frames 585–621.

Circles represent Envisat Image Mode 6 SAR scenes while lines represent SAR interfer-

ograms. Scenes denoted by gray circles are heavily contaminated with topographically

correlated atmospheric signals and were not used. Solid black lines denote interfero-

grams used to invert for fault parameters. Solid gray lines denote coseismic interfero-

grams which were not used. Vertical dashed lines mark the times of the 28–29 Oct. 2008

mainshocks and 9 Dec. 2008 aftershock discussed in the text. . . . . . . . . . . . . . . . . 109

xxvii

6.5 Interseismic velocities, coseismic offsets and residuals for the 28–29 Oct. 2008 earthquakes

and the 9 Dec. 2008 earthquake. A.) Interseismic velocities relative to the stable Indian

Plate. Thick black lines without arrows represent regional faults (see Figure 6.1). B.) Co-

seismic offsets and C.) residuals from the 28–29 Oct. 2008 earthquake. Displacements for

stations KHST and SHRG are poorly defined due to low number of post-seismic observa-

tions. Stations ZART and CHTRwere established in 2009 and therefore have no pre-seismic

position measurements. Black lines represent the rupture planes determined from in-

version of InSAR data. D.) Coseismic displacements and E.) residuals for the 9 Dec. 2008

earthquake. The proximity of station KACH to the epicenter combined with fortunate

post-seismic occupation timing makes this the only station for which I am able to esti-

mate displacements. Black lines represent rupture determined from inversion of InSAR

data. The error ellipses represent formal uncertainties for the coseismic displacements

as measured from the time series for each station and certainly represent a best case sce-

nario. The residual displacements are calculated by removing the best-fitting coseismic

model determined from inversion of InSAR data. . . . . . . . . . . . . . . . . . . . . . . . 111

6.6 Epicentral locations for the 24 Aug. 1931 Sharigh earthquake and the 28 and 29 Oct. 2008

Ziarat earthquakes determined from shaking intensity data. Locations are determined

using the methodology outlined in Chapter 3. The contours represent the 50% and 67%

confidence contours for epicentral location calculated using parameters listed in Bakun

(1999). In each subfigure, filled circles indicate the locations of felt reports, the star in-

dicates the instrumentally determined epicenter and the center of the innermost con-

tour represents the preferredmacroseismic estimate of epicenter. Intensity data are from

Martin and Szeliga (2010). A.) Epicenter of the 24 Aug. 1931 Sharigh earthquake as deter-

mined frommacroseismic data. B.) Epicenter of the 28 Oct. 2008 earthquake as determined

from macroseismic data. C.) Epicenter of the 29 Oct. 2008 earthquake as determined from

macroseismic data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

xxviii

6.7 Graphical comparision of moment tensor solutions from inversion of teleseismic body-

wave data, the Global CMT (Dziewonski et al., 1981) and inversion of InSAR data. Each inver-

sion method is sensitive to deformation in different frequency bands. To illustrate this,

moment tensors are arranged, from left to right, in order of sensitivity to decreasing fre-

quencies (increasing periods) of radiated energy. In cases where more than one subevent

is inverted for, the moment tensor for the subevent with the largest contribution to the

total moment is shown. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

6.8 Envisat interferogram of scenes from 2 Dec. 2008 and 6 Jan. 2009. One fringe corresponds

to 28 mm of change in range. Solid arrow indicates the flight direction of the satellite

and outlined arrow denotes the look direction of the satellite. A.) Original interferogram.

B.) Preferred coseismic elastic dislocation model. C.) Interferogram with coseismic model

removed. Black line denotes the surface projectionof theup-dip edge of the fault identified

from inversion of A. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

6.9 Envisat interferogram of scenes from 6 May 2008 and 2 Dec. 2008. One fringe corresponds

to 28 mm of change in range. Solid arrow indicates the flight direction of the satellite and

outlined arrow denotes the look direction of the satellite. A.) Original interferogram. B.)

Preferred coseismic elastic dislocation model. C.) Interferogram with coseismic model re-

moved. Black lines denotes the surface projection of the up-dip edge of the fault identified

from inversion of A. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

xxix

6.10 Lower hemisphere projection of the moment tensors from the inversion of teleseismic

body-waves for the 9 Dec. 2008 aftershock. Fault plane information for each subevent are

listed in the header as event number, strike, dip, and rake in degrees, depth in km andmo-

ment inN-m. Seismic stationnames are printed vertically and to the left of eachwaveform.

Seismic station locations on the focal sphere are denoted by upper-case letters and corre-

spond to the letter indicated between the station name and the waveform trace. Upper

plot shows P-wave focal sphere and waveforms, while the lower plot shows SH-wave fo-

cal sphere and waveforms. Amplitudes have been normalized to highlight the agreement

between the data (solid line) and the synthetic waveforms (dashed line). The source-time

function along with the time scale for each waveform is shown beneath the P-wave data

for station KMBO. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

6.11 Revised double-difference earthquake relocations for all events in the region during the

periods Feb. 1997–Mar. 1997 and Oct. 2008–Jan. 2009. Earthquakes during this time period

were relocated using phase data from the USGS monthly PDE using the double difference

method of Waldhauser and Ellsworth (2000). Double difference locations for the 9 Dec. 2008

Mw 5.7 aftershock were then compared with the location derived from inversion of the

interferogram in Figure 6.8 to obtain a shift parameter. Revised double difference epicen-

ters were then obtained by applying this shift parameter to all of the double differenced

earthquakes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

xxx

6.12 Lower hemisphere projection of the moment tensors from the inversion of teleseismic

body-waves for the 28 Oct. 2008 aftershock. Fault plane information for each subevent

are listed in the header as event number, strike, dip, and rake in degrees, depth in km

and moment in N-m. Seismic station names are printed vertically and to the left of each

waveform. Seismic station locations on the focal sphere are denoted by upper-case letters

and correspond to the letter indicated between the station name and the waveform trace.

Upper plot shows P-wave focal sphere andwaveforms, while the lower plot shows SH-wave

focal sphere andwaveforms. Amplitudeshavebeennormalized tohighlight the agreement

between the data (solid line) and the synthetic waveforms (dashed line). The source-time

function along with the time scale for each waveform is shown beneath the P-wave data

for station RER. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

6.13 Lower hemisphere projection of the moment tensors from the inversion of teleseismic

body-waves for the 29 Oct. 2008 aftershock. Fault plane information for each subevent

are listed in the header as event number, strike, dip, and rake in degrees, depth in km

and moment in N-m. Seismic station names are printed vertically and to the left of each

waveform. Seismic station locations on the focal sphere are denoted by upper-case letters

and correspond to the letter indicated between the station name and the waveform trace.

Upper plot shows P-wave focal sphere andwaveforms, while the lower plot shows SH-wave

focal sphere andwaveforms. Amplitudeshavebeennormalized tohighlight the agreement

between the data (solid line) and the synthetic waveforms (dashed line). The source-time

function along with the time scale for each waveform is shown beneath the P-wave data

for station DGAR. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

6.14 Photograph of the rupture zone of the 29 Oct. 2008 Ziarat Valley earthquake courtesy of

Din Mohammad Kakar. View looking south into the Kan Tangai (Stone Gorge) from the

village of Wam. No surface rupture was observed in the gorge, but numerous N-S trending

surface cracks were apparent along the roads and hill-slopes. . . . . . . . . . . . . . . . . 122

xxxi

6.15 ERS-1 interferogram of scenes from 13 Sep. 1993 and 22 Nov. 1993. One fringe corresponds

to 28mmof change in range. A.) Original interferogram. Black circle indicates the location

of anMb 4.2 that occurred on 30 Oct. 1993 along the northernmost trace of the Ghazaband

Fault. Solid arrow indicates the flight direction of the satellite and outlined arrow denotes

the look direction of the satellite. B.) Preferred coseismic elastic dislocation model. C.)

Interferogram with coseismic model removed. Black line denotes the surface projection

of the up-dip edge of the fault identified from inversion of A. . . . . . . . . . . . . . . . . 124

6.16 A map of Coulomb stress for a receiver fault with the same geometry as the down-dip

extension of the Deghat-Bannh thrust fault system (gray rectangle). Contours are 50 kPa.

A.) Thrust orientation for the 1931 Sharigh earthquake. B.) Dextral orientation for the 1931

Sharigh earthquake. C.) Sinistral orientation for the 1931 Sharigh earthquake. . . . . . . . 126

6.17 Landsat 7 image from 3 Apr. 2001. Black lines indicated mapped faults, and white lines in-

dicate the surface projection of faults that ruptured during the Oct.–Dec. 2008 earthquake

sequence. Questionmarks are placed to indicatewhere the fault extent is uncertain. Faults

shown with a dot-dash pattern are inferred from inspection of the Landsat image as pos-

sible locations for the 1931 Sharig earthquake. Dark colors along the northern edge of the

image correspond to exposedmafic and ultramafic rocks of theMuslimbagh ophiolite. Let-

ters indicate the location of cities and towns: Pishin (P), Quetta (Q), Sharigh (S), and Ziarat

(Z), star denotes the location of the photograph in Figure 6.14. This image is a combination

of Landsat bands 7, 4 and 2 to highlight differences in lithology. . . . . . . . . . . . . . . . 128

6.18 Idealized shear zone geometry, adapted from Sigmundsson et al. (1995). Black wedges rep-

resent stable boundaries to the shear zone, v is the shear velocity, L is the typical block

length, w is the typical block width, and τ is the rotation rate. . . . . . . . . . . . . . . . 129

xxxii

6.19 Moment release as a function of distance to the Quetta Syntaxis Shear Zone in fault lengths

(25 km). Gray circles indicate earthquakes occurring before 1900 for which moment has

been inferred. Shear zone location is indicated by the vertical dotted lines. Earthquake lo-

cations andmagnitudes are from the EHB Centennial Catalog (Engdahl and Villasenor, 2002),

historical earthquake locations (gray) are from Pakistan Meteorological Department and NOR-

SAR (2007). Note that, besides the lack of locations for aftershocks to the 1931 Sharigh

earthquake, the entire shear zone has ruptured in the past century. . . . . . . . . . . . . 130

7.1 Summarymapof interseismic deformation, as determinedusing space geodetic techniques,

across the western boundary of the Indian Plate. Interseismic deformation rate and fault

locking depth are indicated for three transects (black stippled rectangles) across the plate

boundary and one transect across the western edge of the Sulaiman Lobe (white stippled

rectangle). Thin black lines represent the location of major regional faults, thick black

lines represent the approximate location of fault segments known to have ruptured in

historical times. The years of select major historical earthquakes are shown near the seg-

ments believed to have ruptured. GPS velocities are shown in a Eurasian-Plate-fixed refer-

ence frame (Altamimi et al., 2007); the velocity of TURT is 29.96±0.42 mm/yr. Dark gray

vectors represent observed GPS velocities while light gray vectors represent velocities

predicted by motion about the pole of relative motion between the Indian Plate and the

Eurasian Plate. The map is an oblique Mercator projection about the pole of relative mo-

tion between the Indian and Eurasian Plates, thus points on the stable Indian Plate show

velocity vectors parallel to the lower edge of the figure. Thrust faults are shownwith black

triangles on the hanging wall, all other faults are strike-slip. Fault names are indicated in

Figure 4.1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

xxxiii

7.2 Seismicity and locations of historical fault rupture along the western boundary of the In-

dian Plate. The dashed line represents the boundary of the Katawaz Block (Chapter 4).

A.) Earthquake locations from 1964–2010 from the ISC catalog and moment tensors from

1976–2010 from the Global CMT Project. Filled segments of the moment tensors represent

the compressional quadrants for the best-fitting double-couple. B.) Approximate rupture

lengths for major historical strike-slip earthquakes. Rupture lengths were calculated us-

ing the relationships for strike-slip earthquakes listed in Wells and Coppersmith (1994) us-

ing published estimates of moment magnitude. The 375 km segment between the 1892

Chaman earthquake and the 1505 Kabul earthquake has no known major historical seis-

micity. The 1842 Jalalabad earthquake (Appendix B)was likely a thrust faulting earthquake

and is shown for completeness. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

Chapter 1

Introduction

1.1 The Indian Plate

The collision between the Indian Plate and the Eurasian Plate, which began during the early Ceno-

zoic Era (Molnar and Tapponnier, 1975), is responsible for some of the most significant relief on the planet.

Present day estimates of the convergence rate between these two plates using observations from perma-

nent and campaign Global Positioning System (GPS) stations are close to 38 mm/yr near Hyderbad, India

(Altamimi et al., 2007). Previous studies havemainly focussed on quantifying convergence across the north-

ern boundary between the Indian and Eurasian Plates (Jackson andMcKenzie, 1984; Treloar and Coward, 1991;

Bilham, 2004). Because the Indian Plate is rotating counter-clockwise relative to Eurasia, convergencewith

the Eurasian Plate decreases westwards of Hyderabad, India. The relative velocity along the transform

fault that separates India from Eurasia on the west has never before been measured directly and is the

subject of this Dissertation. This transform boundary includes numerous subsidiary faults and fold-belts

and is known as the Chaman Fault System (Wellman, 1966; Lawrence et al., 1992). The Chaman Fault itself

represents the westernmost tectonic structure in what is a 150–300 km wide belt of diffuse deformation

forming the western boundary between the Indian and Eurasian Plates.

1.2 Geodetic and Seismic Observations

In this thesis, I utilize various geodetic and seismological measurements to quantify and elucidate

tectonic processes along the western boundary of the Indian Plate. Each of these various measurement

techniques provides knowledge of different temporal and spatial aspects of the seismic cycle. In order

2

to examine patterns in seismicity over the longest temporal span, I use non-instrumental analysis tech-

niques, that consist of archival reports of shaking and destruction caused by large earthquakes. With

certain caveats, these macroseismic data may be analyzed to provide coarse resolution of epicentral loca-

tion and magnitude. For data in the past two decades, I use long-period ground velocity data in the form

of seismic waveforms. These waveforms, recorded at seismic stations with global coverage provide more

precise epicentral information thanmacroseismic data and also provide constraints on earthquake source

depth and possible fault rupture planes.

Geodetic data in the form of direct surface height measurements may be obtained using spirit-

leveling. These data, first collected on the Indian subcontinent during the late 19th century, provide

constraints on geoid height along a transect, and, when combined with prior observations, differential

heights. The effort involved in collecting spirit-leveling data with a high temporal resolution results in

infrequent reoccupation of benchmarks with detailed spatial coverage over limited areas (typically < 1

km transects). Differential heights from spirit-leveling provide excellent vertical resolution of coseismic

ground deformation along fortuitously placed survey lines.

Modern estimates of height are routinely measured along with high precision location estimates

using GPS receivers. When continuously operated from a permanent stable benchmark, relative positions

may be obtained to an accuracy of 2–5 mm horizontally and 6–15 mm in height (Segall and Davis, 1997).

While continuously operating GPS stations provide a vast improvement over spirit-leveling in the tem-

poral domain, they suffer in the spatial domain in being essentially point measurements and are often

separated by large distances (20–200 km).

Synthetic Aperature Radar (SAR) data provide dense spatial coverage at the expense of tempo-

ral resolution for geodetic applications. Interferometric SAR (InSAR), produced by differencing two SAR

scenes, provides information about the relative motion of the ground when the SAR scenes are obtained

at different times from similar vantage points. Deformation data obtained by any single InSAR image

provides spatially dense measurements of only one component of deformation, and can limit the capa-

bility of using InSAR alone to fully describe a deformation field (Burgmann et al., 2000). However, this

high spatial density of deformation information at low temporal sampling complements GPS positioning

3

data. Although it is possible to translatemeasured phase differences from an InSAR image tommprecision

line-of-sight measurements, the relationship between phasemeasurement and ground deformation is de-

pendent on the atmospheric conditions along the line-of-sight during the acquisition of each SAR scene

(Massonnet and Feigl, 1995). In Baluchistan, the arid climate and overall absence of vegetation provides

ideal conditions for InSAR analysis.

1.3 Thesis Outline

In Chapter 2, I begin by analyzing felt reports of seismic shaking across the Indian plate to deter-

mine the gross properties associated with the attenuation of ground shaking with distance. These data

are compared to results from previous studies on other tectonic plates to place the properties of the In-

dian Plate in a global context. With knowledge of the behavior of the attenuation of seismic waves as

observed through popular reporting, I then analyze a new and voluminous catalog of historical shaking

records from the Indian Plate (compiled bymy colleague S. Martin) to examine the limits of macroseismic

data on the derived quantities I seek, epicentral location and magnitude. In Chapter 2, I discuss historical

earthquakes on the Indian Plate and their felt reports. Eight thousand three hundred and thirty nine in-

tensity observations have been evaluated for earthquakes that occurred on the Indian subcontinent and

surrounding plate boundaries from the 17th century to the present. They characterize 570 earthquakes,

more than 90% of which occurred in the past two centuries. I summarize these data graphically in the

form of a spatially averaged intensity map for the subcontinent, a map that emphasizes the features of

many previously published earthquake hazard maps for the Indian plate, but which more faithfully de-

picts regional amplification and attenuation. I also estimate the probable return time for future damaging

shaking in five of India’s largest cities.

In Chapter 3, I utilize this catalog of intensity data to quantify uncertainties in the location andmag-

nitude of historical seismicity on the Indian subcontinent. This comprehensive, consistently interpreted

new catalog of felt intensities for India includes intensities for 570 earthquakes, of which, instrumental

magnitudes and locations are available for 100. I use the intensity values for 29 of these instrumentally

recorded events to develop new intensity versus attenuation relations for the Indian subcontinent and the

4

Himalayan region. I then use these relations to determine the locations and magnitudes of 234 historical

events using the method of Bakun and Wentworth (1997). For the remaining 336 events, intensity distribu-

tions are too sparse to determine magnitude or location. I evaluate magnitude and epicentral accuracy

of newly located events by comparing instrumentally-derived with intensity-derived locations for 29 cal-

ibration events for which more than 15 intensity observations are available. With few exceptions, most

intensity-derived locations lie within a fault length of the instrumentally determined location. For events

where the azimuthal distribution of intensities is limited, I conclude that the formal error bounds from

the regression of Bakun and Wentworth (1997) do not reflect the true uncertainties. Specifically, I also find

that the regression underestimates the uncertainties of the location andmagnitude of the 1819 Allah Bund

earthquake, for which a location has been inferred frommapped surface deformation. Comparing my in-

ferred attenuation relations to those developed for other regions, I find that attenuation for Himalayan

events is comparable to intensity attenuation observed in California (Bakun andWentworth, 1997), while in-

tensity attenuation for cratonic events is higher than intensity attenuation reported for central/eastern

North America (Bakun et al., 2003). Further, I present evidence that intensities of intraplate earthquakes

have a non-linear dependence on magnitude, such that attenuation relations based largely on small-to-

moderate earthquakes may significantly overestimate the magnitudes of historical earthquakes.

In Chapter 4, I discuss the strain accumulation rate along the western boundary of the Indian Plate

utilizing both InSAR and GPS measurements. The western boundary is defined by the Chaman Fault Sys-

tem, the on-land transform separating the Indian and Eurasian plates. From the Arabia/Eurasia/India

triple junction offshore of the Makran coast the Chaman Fault System passes north through Baluchistan

and trends NNE into Afghanistan before merging with the Himalayan arc in northern Afghanistan. Geo-

logical and plate closure estimates suggest sinistral slip across the Chaman Fault System of between 19 and

35mm/yr over the last 25 Ma. Along the southernmost on-land sections of the fault system near the town

of Las Bela, Pakistan, campaign GPSmeasurements indicate sinistral slip at a rate of nearly 15mm/yr with

a shallow locking depth. Farther north near the town of Chaman, Pakistan, campaign and continuous GPS

measurements indicate that the Chaman Fault is shallowly locked or possibly creeping at the surface at

a rate of 7.5 mm/yr. Immediately north of the town of Chaman, the trend of the Chaman Fault becomes

5

NNE-SSW and enters a transpressional bend. Estimates of interseismic strain accumulation rates from

InSAR analyses of this segment of the Chaman Fault indicate that the fault is also shallowly locked and

accumulating strain at a rate of 16.8 mm/yr. The modern prevalence of shallow locking depths along the

length of the Chaman Fault System between the Makran Coast and the Ghazni Province of Afghanistan

suggests that large strike-slip earthquakes (Mw > 7) typical of continental scale transform boundaries

are unlikely on the Chaman Fault.

With this understanding, I focus on a destructive sequence of earthquakes along thewestern bound-

ary between the Indian Plate and the Eurasian Plate during the 1930’s. This sequence began with a large

Mw 6.8 strike-slip earthquake in the Quetta Syntaxis shear zone and culminated in aMw > 7.5 strike-slip

event along one of the plate bounding faults killing > 35, 000 people in the city of Quetta, Pakistan. In

Chapter 5, I discuss one of these earthquakes, theMw 7.1Mach earthquake. Surface deformation associ-

ated with theMw 7.1 27 Aug. 1931 earthquake near Mach in Baluchistan is quantified from spirit-leveling

data and detailed structural sections of the region interpreted from seismic reflection data constrained

by numerous well logs. Mean slip on the west-dipping Dezghat-Bannh fault system amounted to 1.2 m

on a 42 km × 72 km thrust plane with slip locally attaining 3.2 m up-dip of an inferred locking line at

approximately 9 km depth. Slip also occurred at depths below the interseismic locking line. In contrast,

negligible slip occurred in the 4 km near the interseismic locking line. The absence of slip here in the

4 years following the earthquake suggests that elastic energy there must either dissipate slowly in the

interseismic cycle, or that a slip deficit remains, pending its release in a large future earthquake. Elastic

models of the earthquake cycle in this fold-and-thrust belt suggest that slip on the frontal thrust fault is

reduced by a factor of 2 to 8 compared to that anticipated from convergence of the hinterland, a partition-

ing process that is presumably responsible for thickening of the fold-and-thrust belt at the expense of slip

on the frontal thrust. Near the latitude of Quetta, campaign GPSmeasurements indicate that convergence

is approximately 5 mm/yr. Hence the minimum renewal time between earthquakes with 1.2 mmean dis-

placement should be as little as 240 years. However, when the partitioning of fold-belt-convergence to

frontal-thrust-slip is taken into account the minimum renewal time may exceed 2000 years.

Finally, in Chapter 6, I discuss a recent earthquake sequence whose deformation was observed us-

6

ing multiple space-based geodetic methods. In October 2008, two Mw 6.4 earthquakes occurred within

11 hours and 15 km of each other, 40 km NE of Quetta in northern Baluchistan, causing significant dam-

age throughout the Ziarat Valley. Initial interpretations suggested that the earthquakes had occurred on

contiguous segments of a shallow NW-SE trending dextral fault in spite of the absence of mapped faults

with this trend. A relocation of the mainshocks and aftershocks using a double-difference methodology

was confirmed using the surface deformation field of a large aftershock imaged by InSAR. The relocated

mainshocks were subsequently used to interpret InSAR imagery significantly decorrelated by landsliding

in the epicentral region, revealing that the two mainshocks had occurred on parallel NE-trending sinis-

tral faults. A reinterpretation of historical earthquakes near the 2008 earthquake sequence suggests that

“book-shelf” faulting extends to the NW and SE of the Ziarat valley and accommodates overall dextral

shear arising from the southward advance of the Sulaiman Lobe past the eastward advance of the north-

ern Kirthar ranges. Regional GPS data suggest that an effective dextral slip rate of approximately 17.0

mm/yr is accommodated by a previously unrecognized system of at least 5 approximately 25 km long,

NE-trending sinistral faults spaced approximately 15 km apart.

Chapter 2

A Catalog of Felt Intensity Data for 570 Earthquakes in India from 1636 to 2009

2.1 Introduction

Catalogs of historical Indian earthquakes occurring in the past 450 years contain errors in date,

location and magnitude, and list few intensity data in a form suited to numerical analysis. The following

account addresses this deficiency by presenting a unified analysis of intensity data assessed from accounts

of damage, or from felt perceptions of earthquakes. As such it omits some earthquakes for which no

intensity data are available. In contrast it includes several earthquakes missing from previous catalogs

. With few exceptions, the listing in based on original source materials archived in Indian and European

libraries, regional newspapers, private letters and diaries, and government reports. For earthquakes later

than 2000, eyewitness accounts provided via the World Wide Web or communicated in person have also

been included. In total, 570 earthquakes are listed using 8339 intensity evaluations based on the European

Macroseismic Scale (EMS-98) (Grunthal and Levret, 2001). Of these earthquakes, 7 occurred before 1800, 240

occurred between 1800–1900, 158 occurred between 1900–1960 and a further 165 earthquakes occurred in

the period 1960–2009 (Figure 2.1). The data were obtained from the electronic supplement to Martin and

Szeliga (2010) and list the latitude, longitude and location of each felt report and its inferred intensity. The

formats for the tables in the electronic supplement are shown in Tables 2.1 and 2.2.

Early earthquake catalogs for India consist of anecdotal information, dates and locations, but few

data suited to quantitative evaluation of intensity (Baird Smith, 1843a,b, 1844; Mallet and Mallet, 1858; Old-

ham, 1883; deMontessus deBallore, 1896). Events listed in early catalogswere often repeated in later catalogs,

and these events were subsequently included inmore recent compilations (Bapat et al., 1983; Srivastava and

8

100

200

300

400

500

600

Cum

ula

tive N

um

ber

of E

art

hquakes

1600 1650 1700 1750 1800 1850 1900 1950 2000

Year

1

2

5

10

20

50

100

200

500

1000

Inte

nsity O

bserv

ations p

er

Eart

hquake

Figure 2.1: A cumulative histogram of earthquakes per 50 year period in the historical seismic catalog(right hand axis). Vertical bars topped with circles (left hand axis) show observations per earthquake.

9

Table 2.1: The first dozen earthquakes from the electronic supplement toMartin and Szeliga (2010) to illus-trate format. Columns Year, Month and Day refer to the date of an event in local time. For earthquakeswith more than seven intensity observations (column Number of Observations), the approximate epi-central location is listed (columns Longitude, Latitude). The number of observations corresponds to thenumber of intensity reports listed in the electronic supplement toMartin and Szeliga (2010). A geographicregion designator is defined for some events (column Earthquake). This column serves as a referencecolumn to groups of intensity observations in Table 2.2

NumberDate Longitude Latitude of Earthquake

Observations1636-08-29 11669-06-23 11676-08-26 11736-03-24 11762-04-02 22.4 92.2 9 CHITTAGONG-17621779-??-?? 11784-12-?? 11800-10-19 11803-09-01 30.7 78.8 25 BARAHAT-18031808-06-04 11810-04-01 11810-05-13 1

Table 2.2: The first 5 earthquakes of 570 from the electronic supplement to Martin and Szeliga (2010).Columns Year, Month and Day refer to the date of an event in local time. Columns Longitude and Lati-tude refer to the location of the intensity observation. Column EMS-98 lists assessed EMS-98 intensities(Grunthal and Levret, 2001). The geographic location of each observation is listed in column Location. Col-umn Earthquake serves to group observations from the same earthquake and refers to the geographiclocation of each earthquake in Table 2.1. Earthquakes with fewer than 2 observations are not assignedgeographic locations.

Date Longitude Latitude EMS-98 Location Earthquake1636-08-29 72.81 21.19 3 Surat1669-06-23 74.79 34.08 5 Srinagar1676-08-26 86.94 21.48 4 Balasore1736-03-24 74.79 34.08 7 Srinagar1762-04-02 88.35 22.57 3 Calcutta CHITTAGONG-17621762-04-02 88.386 22.88 3 Chandernagore CHITTAGONG-17621762-04-02 91.838 22.342 5 Goyparah CHITTAGONG-17621762-04-02 91.665 22.552 6 Akulpoor-Bansbaria CHITTAGONG-17621762-04-02 91.773 22.297 6 Howla CHITTAGONG-17621762-04-02 91.826 22.349 7 Chittagong/Islamabad CHITTAGONG-17621762-04-02 92.101 22.133 7 Dahrampoor CHITTAGONG-17621762-04-02 92.065 22.168 7 Do Hazari CHITTAGONG-17621762-04-02 92.084 22.367 8 Bahngoo Changee CHITTAGONG-1762

10

Ramachandran, 1985; Ramachandran and Srivastava, 1991) supplemented by new earthquakes, and by newly

discovered archival information. Hence, these new catalogs include many erroneous entries from earlier

catalogs. The global Catalog of Significant Earthquakes by Dunbar et al. (1992) lists all these earthquakes

uncritically making it impossible to judge which accounts should be rejected. As Ambraseys (1971) noted,

the repetition of error is common to many catalogs of earthquakes that have not been evaluated from

primary source materials.

The present account includes data assessed from primary sources, or from sources that reproduce

the raw data from which intensity may be evaluated or verified (Martin and Szeliga, 2010, electronic sup-

plement). I emphasize that the list of the locations of Indian earthquakes in Martin and Szeliga (2010) is

subordinate to the listing of perceived and felt observations of intensity (electronic supplement toMartin

and Szeliga (2010)) because the determination of epicentral location is subject to interpretation. The tabu-

lated intensity data are quantified from reports at locations that are rarely at the epicenter. Therefore, the

location and magnitude of all pre-instrumental earthquakes in India derived from these data are uncer-

tain except in those rare locations where surface deformation has been recorded (e.g. 1819 Kachchh and

1897 Shillong). The entries are listed in chronological order. In Chapter 3 I calculate epicentral locations

for many of these earthquakes using the methods of Bakun and Wentworth (1997).

Beginning in the late 1800’s, the Geological Survey of India and other agencies compiled studies

of significant earthquakes. In many official government reports, a simplified description of the building

stock considered characteristic of a whole village is used (e.g. 1967 Koyna earthquake (Tandon and Chaud-

hury, 1968)). In some cases this generalized description of the building stock is classified into Types A, B

and C as defined in Grunthal and Levret (2001). However, many of these reports omit descriptions of shak-

ing experienced by people. From these government reports, descriptions of damage, in some instances

accompanied by photographs, were used to evaluate intensities. Of the 43 events from Ambraseys and Dou-

glas (2004) 28 have been re-evaluatedwhere it was possible to locate first-hand accounts or official reports.

None of the listed intensities have been repeated from maps or previously published listings. Where au-

thentic primary source materials are unavailable for a particular earthquake, those accounts have been

excluded from the final listing.

11

2.2 Intensity Scale

In previous studies, various intensity scales were used to evaluate earthquakes in India. Oldham

(1899) notes that early European scales listed inappropriate criteria for the assessment of acceleration-

related damage to indigenous structures, and for the 1897 Assam earthquake he chose to use his own

simplified scale rather than the thenprevalent Rossi-Forrel scale. Later studies of earthquakes adopted the

Modified Mercalli scale (e.g.Middlemiss, 1910) or the MSK-64 scale (Medvedev et al., 1965). Intensities in the

present study use the European Macroseismic Scale (EMS-98) (Grunthal and Levret, 2001, see also Appendix

A), a successor to the MSK-64 intensity scale. I note that MSK-64 listings of Ambraseys and Douglas (2004)

are numerically indistinguishable from the EMS-98 evaluations for those accounts I have compared. As in

Ambraseys andDouglas (2004) assessment of local intensities avoids observations based on, or contaminated

by, ground deformation, landslides, liquefaction, seismic seiches and surface faulting. Numerous accounts

that fall into these categories have thus been excluded from the catalog.

Many of the intensities evaluated lie in the intensity range II–V. These are differentiated from

sparse data as follows: reports that stated an earthquake was “barely felt” or “very slight” were assigned

intensity II while those that stated an earthquake was “slight” or “mild” were assigned intensity III. Re-

ports that spoke of tremulous motion, rumbling sounds, etc. were assigned intensity IV. Grade I damage

(Grunthal and Levret, 2001) to structures begins at intensity V with the appearance of structural cracks,

tiles and plaster being dislodged etc. Above intensity V, the following critera are used: Masonry damage

begins at intensity VI, and accounts of this level of damage often estimate the numbers of buildings af-

fected i.e. “a few”, “many”, or “most”, which together with human perceptions, permit us to distinguish

between intensities VI and VII. Photographs, if available, were used only to supplement intensity assign-

ment. I further note that photographic evidence is often biased towards the most damaged structures,

since undamaged structures are rarely photographed (Hough and Pande, 2007). Because of this known bias,

photographs were never used solely to determine intensities.

12

2.3 Reporting Consistency and Completeness

Earthquakes in India, as elsewhere, result in felt reports where the density of reporting is propor-

tional to the density of population. The number of felt reports is further dependent on the propensity

of a population to commit their perception of shaking or perceived damage to some form of permanent

record. Large urban centers contain a range of vulnerable structures, with people of different levels of

awareness, and the record of their perceptions depends much on the prevailing traditions of personal

diaries and responsibilities of the press and government offices to print these materials. Not only have

these reporting habits changed throughout the past few hundred year, but so have the style of buildings

and construction materials used to make these buildings.

Prior to the 18th century, reporting was sparse and mainly undertaken by official historians and

intellectuals. By the late 19th century the reporting of earthquakes by scattered colonial observers became

more verbose and eloquent. During the 20th century, seismologists began proactively collecting intensity

data and initiating studies of specific earthquakes. The mid 20th century is characterized by a decline in

the number of people writing and presenting personal diaries or sending notes to newspapers. Instead, we

must trust the record of professional reporters trained to gather and print information in the local and

national media. In the past decade the internet has given many people the opportunity to report their

perceptions rapidly. Specific blackouts in reporting have also occurred, such as during the Second World

War, when damage to some cities was classified.

Although reporting improves considerably after 1800, many areas are not represented well, even

at the present time. Thus, it is certain that uneveness in reporting prevails during the time spanned

by the catalog. This is partly because the density of people reporting earthquakes varies spatially, and

partly because public interest in reporting felt intensities has varied significantly with time. Many small

earthquakes may be noted by people but not recorded in news media or public reports. Thus I anticipate

that additional earthquakes and accounts of existing earthquakes will surface in future years that will

supplement the recorded observations listed.

I emphasize that the present catalog is not a complete list of all Indian earthquakes. I estimate that

13

only for M > 8 is the list complete for the Indian subcontinent since 1800. In Chapter 3 I calculate the

Gutenberg-Richter b-value for magnitudes estimated from the intensity data listed in Martin and Szeliga

(2010). The b-value thus determined is approximately 0.3 (compared to instrumental catalogs where the

b-value is ≈ 1.0). This suggests that substantially more than half of all earthquakes M < 6 are missing.

I note, however, that the earthquakes recorded by people are those where populations are dense and

have steadily increased in the past few hundred years. The resulting catalog is thus of intrinsic utility for

estimating seismic hazards to these present large populations.

2.4 Summary of Results

The spatial coverage of intensity observations for India is plotted in Figure 2.2(a). Regions with low

population density, such as the Rajasthan desert, parts of Baluchistan, the Nepal and Assam Himalaya and

the Indo-Burman ranges are sparsely sampled. In contrast, trade and communication routes are manifest

and appear as strings of observations across otherwise uninhabited regions. At many points in Figure

2.2(a) multiple estimates of shaking intensity are available, both from individual earthquakes and from

multiple earthquakes.

From these raw data, I have prepared maps that show the maximum felt intensities at every point

where a felt report has been obtained (Figure 2.3(a)) (Quittmeyer et al., 1979). In regions where the popu-

lation is sparse, the points so obtained are often from isolated accounts. In contrast, in regions of dense

population the larger sample size results in a broader spectrum of observed shaking intensity. To account

for uncertainties in named felt locations, I group all intensity data within a 10 km radius and calculate the

maximum shaking intensity observed in each grouping (Figure 2.3(a)).

Various forms of spatial averaging are possible to make it easier to form general conclusions, and

to suppress extreme values that may be caused by anomalous observations. I choose to interpolate the

grouped data set using a nearest-neighbor scheme with a 50 km search radius (Figure 2.3(b) upper right).

Although amore thorough statistical treatment (e.g. Kozuch, 1995; Bozkurt et al., 2007) would require

even greater sampling in time, for several cities with large and growing populations, sufficient intensity

data are available to begin to form a statistical view of past, and possibly, future shaking. The five largest

14

64˚ 68˚ 72˚ 76˚ 80˚ 84˚ 88˚ 92˚ 96˚

12˚

16˚

20˚

24˚

28˚

32˚

36˚

500 km

(a) Intensity Observations 1636–200964˚ 68˚ 72˚ 76˚ 80˚ 84˚ 88˚ 92˚ 96˚

12˚

16˚

20˚

24˚

28˚

32˚

36˚

500 km

(b) Epicenters 1636–2009

Figure 2.2: (a) Circles indicate the locations of intensity data listed in the electonic supplement to Martinand Szeliga (2010). Regions with low population density, such as the Rajasthan desert, parts of Baluchistan,the Nepal and Assam Himalaya and the Indo-Burman ranges are poorly represented historically. Commu-nication routes and rail lines show up as faint lines in the data. (b) Epicenters for historic earthquakeslisted in the electronic supplement toMartin and Szeliga (2010) determined using the method of Bakun andWentworth (1997).

15

68˚ 72˚ 76˚ 80˚ 84˚ 88˚ 92˚ 96˚

12˚

16˚

20˚

24˚

28˚

32˚

36˚

500 km

(a)68˚ 72˚ 76˚ 80˚ 84˚ 88˚ 92˚ 96˚

12˚

16˚

20˚

24˚

28˚

32˚

36˚

500 km

(b)

68˚ 70˚ 72˚ 74˚

20˚

22˚

24˚

100 km

(c)68˚ 70˚ 72˚ 74˚

20˚

22˚

24˚

(d)

88˚ 90˚ 92˚ 94˚

22˚

24˚

26˚

100 km

(e)

88˚ 90˚ 92˚ 94˚

22˚

24˚

26˚

(f)

Figure 2.3: (a). Maximum shaking intensity observed during the period 1636–2009. (b). Interpolatedmaximum shaking intensity observed during the period 1636–2009. (c). Interpolated maximum shakingintensity in Gujarat. (d). Map of average shear wave velocity down to 30 m (Vs30) for the Indian stateof Gujarat. (e). Interpolated maximum shaking intensity in northeast India. (f). Vs30 map of the north-eastern India. In producing interpolated maximum shaking intensity maps, locations within 10 km of oneanother were binned to account for differences in location names and centers of population over time.Maximum shaking intensity data were interpolated using a nearest neighbor schema. Vs30 maps werederived from 30 arcsecond SRTM V 2.0 data (Farr et al., 2007) using the techniques outlined in Wald andAllen (2007).

16

modern cities in India, Mumbai, Delhi, Bangalore, Kolkata, and Chennai have been shaken numerous times

in the past 200 years by earthquakes. Figure 2.4(a) illustrates maximum shaking as a cumulative number

of observations per year experienced in each major city and Figure 2.4(b) shows the frequency of shaking

at different intensities. Figure 2.4(b) reveals well-behaved curves fromwhich it is possible to conclude the

probability for shaking in a given time window. Although the intensity data for these curves include both

infrequent large and distant earthquakes, and more frequent small but closer earthquakes, the return

times are probably reliable estimates of future shaking. That is, the infrequent larger earthquakes do not

substantially bias the statistics to shorter return times, because there are fewer of them.

The projection of the curves in Figure 2.4(b) to larger intensities than those recorded in the past

200 years in each city is possible, but the predictions are of uncertain reliability. The data in Figure 2.4(b)

follow a function of the form,

log(N) = a+ b(I − 2)

where N is the cumulative number of observations per year for each EMS-98 intensity value I , and a

and b are to be determined. The results of regressing the data to this function are shown in Table 2.3.

Certain of India’s largest cities report shaking (intensity II) more frequently than others. The “a” values

for Delhi and Kolkata are 50% greater than those for Chennai and Bangalore. This is partly due to their

tectonic setting, with cities that are far fromplate boundaries, like Bangalore, showing the longest interval

between shaking at any intensity. Cities closer to plate boundaries, like Delhi (The Himalaya) or Kolkata

(The Indo-Burman Ranges) show the shortest intervals between shaking at a given intensity. Intensity

V shaking in these cities occurs approximately every 15 years. Intensity VII shaking, where well-built

structures begin to show damage, has a forecast return time of approximately 30 years in major cities

such as Delhi and Kolkata, an interval of time comparable to the design life of most structures.

I recognize that the data in Figure 2.4(b) show evidence for incompleteness at both high and low

intensity values. The lowest EMS-98 intensity (I) is, in effect, a “not-felt” observation, and as such, is

expected to be underrepresented in any data set. The number of earthquakes observed in each city con-

sidered is constant over the past 200 years (Figure 2.4(a)). However, I know of no earthquake in India or

17

its surroundings that, in the past 500 years, has repeated. No fault segment has re-ruptured in this time,

with the exception of the eastern plate boundary. Hence, 200 years is a short time interval compared

to the recurrence interval for earthquakes in India. I therefore recognize that high intensity shaking is

undersampled in my data.

Table 2.3: Regression coefficients and anticipated mean return time in years for shaking at EMS-98 inten-sities V, VI and VII for the five largest cities in India.

Return TimeCity a b (Years) for Intensity

V VI VIIMumbai -0.81 -0.27 42 78 145Delhi -0.66 -0.18 16 24 36

Bangalore -1.07 -0.28 81 155 295Kolkata -0.72 -0.14 14 19 26Chennai -1.04 -0.20 44 69 110

2.5 Discussion

My intensity data sample fewer than four centuries of earthquakes and are largely populated by

earthquakes from the past 200 years. I know of no moderate or large earthquake that has repeated in this

time period, even at India’s plate boundaries where crustal deformation rates are at their highest. Hence

an important conclusion is that Figure 2.3 represents an incomplete view of anticipated future shaking.

It is only necessary to reflect that had the Koyna, Killari or Jabalpur earthquakes not occurred in the past

half century the view of shaking in central India would be very different. In that recurrence interval for

earthquakes near the boundaries of the Indian plate are shorter, the intensity maps are more reliable in

these regions than those constructed within central India. It is improbable that the rate of occurrence of

earthquakes prevailing in central India will provide sufficient shaking data to provide reliable maximum

intensity maps for many hundreds of years. For this reason alternative methods to estimate potential

future shakingwill be needed to supplement future hazard studies. Thesemay include the study of surface

and subsurface faults and surface liquefaction features (Rajendran et al., 2008), archaeological and archival

research (Ambraseys and Jackson, 2003; Raghu Kanth and Iyengar, 2006), and the development of physical

18

0

25

50

Cum

ula

tive N

um

ber

of E

art

hquakes

1750 1800 1850 1900 1950 2000

Year

KolkataDelhiMumbaiChennaiBangalore

(a)

−2.00

−1.75

−1.50

−1.25

−1.00

−0.75

−0.50

−0.25

Lo

g(C

um

ula

tive

Nu

mb

er

of

Ob

se

rva

tio

ns p

er

Ye

ar)

1 2 3 4 5 6 7

EMS−98 Intensity

2

5

10

20

50

100

Re

cu

rre

nce

In

terv

al (Y

ea

rs)

MumbaiDelhiBangaloreKolkataChennai

(b)

Figure 2.4: (a) Cumulative number of earthquakes felt in major Indian cities since 1762. (b) Frequency ofmaximum shaking intensities observed in these cities in the past two hundred years. The regression coef-ficients to these data, fit between intensity II and V are shown in Table 2.3. The well behaved form of thesecurves suggests that the probability for future shaking from modest earthquakes can be estimated withreasonable confidence. The estimation of the probable return time of higher intensity shaking from thesecurves is less well constrained. The light gray line is the regression line for Delhi using the coefficientsfrom Table 2.3.

19

models for characterizing stress caused by India’s collision with Asia (Bilham et al., 2003) .

I note that Figure 2.3(b) resembles many previously published seismic hazard maps of the Indian

subcontinent (e.g. from the Global Seismic Hazard Assessment Program http://www.seismo.ethz.ch/

GSHAP/). I caution however, that the maps presented here, based as they are on felt reports, or catalogs

of historical earthquakes, are maps of past shaking rather than future shaking. With this caveat Figure

2.3(b) is potentially superior to previous hazard maps of India in that it represents a spatial average of

intensities that includes the effects of local amplification or attenuation caused by surface properties but

excludes data such as reports of liquifaction and surface faulting. The averaging I impose on this all-India

scale smoothes the details of local amplification of most utility to hazard estimates. In some regions finer

zonation is possible from the data I provide in the electronic supplement toMartin and Szeliga (2010) (Fig-

ures 2.3(c) and 2.3(e)). Figures 2.3(d) and 2.3(f), show the estimated average shear-velocity to 30m derived

from the roughness of 30 arcsecond SRTM version 2.0 data (Farr et al., 2007) using the techniques outlined

inWald and Allen (2007). Thesemaps summarize seismic site conditions and are a proxy for ground-motion

amplification, which are partially reflected in my maximum shaking intensity maps.

2.6 Conclusions

I have used primary sources to assess 8339 macroseismic observations from 570 historical earth-

quakes occurring on the Indian subcontinent using the EMS-98 intensity scale. I have summarized these

data graphically and note similarities between maps of maximum felt intensity and previously published

seismic hazard maps.

Using the maximum observed intensity per earthquake, I have sufficient data to form conclusions

concerning the average time between strong shaking for five large Indian cities. My use of maximum

shaking intensity is biased towards regions in a city where amplification may occur, thus the intervals

between shaking at a given intensity are pessimistically short. The data are insufficiently dense to un-

dertake microzonation within each city. In Delhi and Kolkata, I find that the interval between potentially

damaging shaking (EMS-98 VII) is comparable to the design life of most structures, and should thus be

included in construction codes.

20

Chapter 3 analyses the data presented here in terms of their implications for attenuation of seis-

mic waves traversing the Indian craton and its plate boundaries. In this second article the location and

magnitude of the historical earthquakes discussed here are evaluated from relationships derived between

recent intensity observations and instrumental magnitudes.

Chapter 3

Intensity, Magnitude, Location and Attenuation in India for Felt Earthquakes since 1762

3.1 Introduction

Despite awritten history extendingmore than threemillennia the location andmagnitude of earth-

quakes in the Indian subcontinent and its surroundings prior to 1900 remain largely unquantified. The

catalog presented in Chapter 2 of 8339 felt reports of 570 earthquakes since 1636 permits this shortcoming

to be addressed. More than 98% of the earthquakes in this macroseismic catalog occurred after 1800, and

more than 50% since 1900. In this article I quantify attenuation versus distance relationships for India and

from these I determine the probable magnitudes and locations of earthquakes that occurred before the

instrumental catalog.

Previous studies have undertaken similar investigations using less complete data with variable and

uncertain quality. In 1996, Johnston used published intensity values to derive attenuation parameters for

the Indian subcontinent (Johnston, 1996). However, these intensity values were not consistently deter-

mined, and were biased by the inclusion of observations influenced by liquefaction and by inattention to

the effects of building fragility common to early reports. From these data, Johnston (1996) derived relations

between isoseismal area and earthquake magnitude.

More recently, a number of studies have carefully and systematically reinterpreted availablemacro-

seismic data for a number of important historical earthquakes. Ambraseys and Jackson (2003) present inten-

sity evaluations and approximate magnitudes for several early events in the Himalaya and southern Tibet

(1411, 1505, 1555, 1713, 1751, 1803 and 1806). Ambraseys (2004) assigns intensity values for a Bangladesh

earthquake in 1664 and discusses the location of an earthquake in Sindh in 1668. Ambraseys and Douglas

22

(2004) present re-evaluated intensities from 43 earthquakes in northern India and use inferred felt areas

to estimate attenuation.

Recent events, such as the 2001 Bhuj earthquake, have been the subject of extensive, traditional,

ground-based intensity surveying of damage andother effects (Pande andKayal, 2003). Additionally, Internet-

basedmethods (Wald et al., 1999a, Amateur Seismic Centre (http://www.asc-india.org)) have now begun

to yield objectively determined intensity distributions for moderate and large earthquakes through the

use of standardized questionnaires.

Recent efforts notwithstanding, systematically and carefully determined intensities have remained

lacking for bothmoderatehistorical earthquakes and formostmoderate and large instrumentally recorded

earthquakes in India. This new catalog of felt earthquakes and intensities, compiled from extant records

in colonial libraries and newspaper accounts provides a new, rich source of information for the past two

centuries. Intensity values in this catalog were assessed from the original sources using the European

Macroseismic Scale 1998 (EMS-98) (Grunthal and Levret, 2001). This new catalog includes 234 historical

earthquakes ranging in magnitude from 4 to 8.6, that I judge to have a sufficient number of intensity ob-

servations to permit the evaluation of their epicentral parameters. The results of these evaluations are

listed in electronic supplement.

This important new catalog provides the basis for determining intensity attenuation relations for

India and for determining locations and magnitudes for historical events for which sufficient macroseis-

mic information exists. I conclude my study by discussing examples of four earthquakes from the 19th

century.

3.2 Data and Methods

The intensity values from the catalog used to derive the attenuation relationships for this study

reveal significant scatter at all distances. Although some of this scatter is expected to result from impre-

cision in intensity assignments, (for example where structural fragility cannot be adequately assessed)

rich, objectively determined intensity distributions (e.g. Wald et al., 1999b; Atkinson and Wald, 2007) reveal

that intensities do vary substantially as a consequence of local site geology and other factors. Due to

23

unknown variations in the precise location of repeated observations, the calculation of meaningful site

corrections is not possible. Thus, I do not consider site corrections in this analysis.

Previous studies of intensity attenuation in the Indian subcontinent have used methods based on

the area contained within a contour of specific intensity (e.g. Johnston, 1996; Ambraseys and Douglas, 2004).

These methods assign epicentral locations and magnitudes based on the location of maximum shaking

and the areal extent of isoseismal contours. In this study, I use the method of Joyner and Boore (1993) to

derive intensity attenuation relationships for the Indian subcontinent empirically. The functional form

of the intensity attenuation relationship used in this study is as follows:

I = a+ bMw + cR+ d log(R) (3.1)

where R is the hypocentral distance, Mw is the moment magnitude and a, b, c and d are constants to be

determined. Equation (3.1) is derived by assuming that intensity is logarithmically proportional to the

energy density of a point source (Howell and Schultz, 1975). The cR and d log(R) terms are generally taken

to reflect intrinsic attenuation and geometrical spreading, respectively, although in practice these two

terms are difficult to resolve independently.

A one-stage maximum likelihood methodology is used to derive the intensity attenuation relation-

ship using 29 calibration events (Joyner and Boore, 1993). The calibration events consist of earthquakes

since 1950 withmore than 15 felt intensity reports (Figure 3.1). Although I give preference to earthquakes

with hypocenters in the Centennial catalog (Engdahl and Villasenor, 2002), I utilize other hypocentral cat-

alogs for more recent earthquakes. If an event is not listed in the Centennial Catalog, I use hypocen-

tral estimates from the Bulletin of the International Seismological Centre (ISC) and thereafter, the USGS

NEIC Monthly Hypocenter Data File (MHDF). Preferred moment magnitude estimates are from the Global

Centroid Moment Tensor Project. If an event is not listed in the Global CMT, I use moment magnitude

estimates from the Centennial Catalog, ISC, or the MHDF in decreasing order of preference (see Acknowl-

edgements). For five calibration events (4.1 < M < 5.3), only body wavemagnitude (mb) estimates were

available. Converting these body wave magnitudes to moment magnitudes using a published linear re-

lationship resulted in attenuation relationship coefficients that were statistically indistinguishable from

24

the uncorrected magnitudes. I therefore have chosen to retain the original body wave magnitudes during

inversion. For the largest event in the catalog, the 1950 Chayu earthquake in eastern Assam, I use the

hypocentral location and magnitude listed in Chen and Molnar (1977).

I first use a least squares approach to estimate parameters a–d in equation (3.1) using themagnitude

of all calibration earthquakes as well as the hypocentral distance to each observation. The least squares

inversion is weighted by a covariance matrix that includes off-diagonal terms that account for intra-

earthquake observational variance. The inversion is performed by inverting the normal equations with

the off-diagonal terms in the covariance matrix being determined using a maximum likelihood method-

ology.

Utilizing the attenuation relation derived from the methods outlined above, I then use the method

outlined in Bakun and Wentworth (1997) to determine epicenters and magnitudes. For each earthquake I

create a 5◦ × 5◦ grid of trial hypocenters centered on the instrumentally determined hypocenter with a

grid spacing of 5 arc-minutes. If no instrumental hypocenter is available, I use the geometrical centroid

of all of the intensity observations weighted by their EMS-98 value and a depth of 15 km. For each trial

hypocenter I calculate the slant distance to each intensity observation and solve equation (3.1) forMw. A

weightedmeasure of the dispersion of the magnitude estimates is then calculated at each grid point using

the following equation:

σ =

(∑i(Wi(Mi − M))2∑

iW2i

) 12

(3.2)

with,

Wi =

0.1 + cos( ∆iπ

(2)(150)) ∆i < 150 km

0.1 ∆i > 150 km,

where ∆i is the distance from the trial hypocenter to each intensity observation i, Mi is the magnitude

estimated from equation (3.1) for observation i and M is the mean magnitude at the trial epicenter. I

then choose the trial epicenter that minimizes equation (3.2) as the preferred epicentral estimate and

its associated M as the preferred magnitude estimate. In a scenario where all intensity observations are

25

60˚ 65˚ 70˚ 75˚ 80˚ 85˚ 90˚ 95˚ 100˚

10˚

15˚

20˚

25˚

30˚

35˚

40˚

Figure 3.1: Epicentral locations of 29 calibration events. I have excluded earthquakeswith depths in excessof 40 km. Eventsmarkedwith diamondswere used to determine cratonic attenuationwhile eventsmarkedwith circles were used to determine Himalayan attenuation.

26

given equal weight, (i.e choosingWi = 1.0 for all∆i), equation (3.2) becomes the sample standard devia-

tion. Thus, the trial epicenter that minimizes equation (3.2) will be referred to as the minimum deviation

epicenter.

In general, intensity observations show a rapid decay close to the epicenter; this behavior indicates

that intensity observations near the epicenter aremore sensitive to the epicentral location andmagnitude

than observations farther away. Thus, I choose a function, Wi, that gives greater weight to observations

that are closer to the trial epicenter. While Bakun andWentworth (1997) note that the 150 kmcutoff distance

chosen for the weighting function is arbitrary, I retain this value to facilitate direct comparison of these

results with those of Bakun and Wentworth (1997). A possible benefit of retaining a cutoff distance of 150

km is that it down-weights potentially magnified observations that may result from critically reflected

seismic phases such as SmS. In India, Moho depths vary from greater than 50 km on the Craton (Gupta

et al., 2003) to 40 km beneath the Himalaya (Monsalve et al., 2008). Given a hypocentral depth of 15 km, one

could reasonably expect SmS to first appear between 120–150 km from an epicenter.

3.3 Results

I calculate separate attenuation parameters for earthquakes in the subcontinent (craton) and the

Himalaya, in addition to evaluating the parameters for the entire data set (Table 3.1). Additionally, Figure

(3.2) shows the distribution of intensity data used to calculate the attenuation parameters as a function of

moment magnitude.

Table 3.1: Intensity attenuation relationship parameters for India, the Indian Craton and the Himalaya.Columns a, b, c, and d refer to the variables in equation (3.1).

NumberProvince of a b c d

EventsIndia 29 5.57±0.58 1.06±0.07 -0.0010±0.0004 -3.37±0.25Craton 17 3.67±0.79 1.28±0.10 -0.0017±0.0006 -2.83±0.30

Himalaya 12 6.05±0.94 1.11±0.10 -0.0006±0.0006 -3.91±0.38

To investigate the self-consistency of my results, I utilize a cross validation scheme (Efron and Tib-

shirani, 1994) to characterize the predictive ability of this data set. I determine attenuation relationships

27

3

4

5

6

7

8

9

Mom

ent M

agnitude

1 10 100 1000

Distance to Centroid (km)

Assam 1951

Kashmir 2005

Himalaya EMS−98 data

(a)

3

4

5

6

7

8

9

Mom

ent M

agnitude

1 10 100 1000

Distance to Centroid (km)

Bhuj 2001

Craton EMS−98 data

(b)

Figure 3.2: Intensity distributions for the data used to calculate the attenuation parameters in Table (3.1).(a) Distance to earthquake centroid versus moment magnitude for events in the Himalaya. (b) Distance toearthquake centroid versus moment magnitude for events on the Craton.

using subsets of 21 instrumentally recorded calibration events randomly chosen without replacement

from the original list of 29 calibration events. I then use the resulting attenuation relationship to deter-

mine the locations and magnitudes of the remaining 8 calibration events. This procedure is repeated to

create 100 cross-validation samples. The resulting statistics show a median epicentral misfit of 53 km and

a magnitude misfit of 0.38Mw.

3.3.1 Comparisons with previous attenuation studies

As noted, previous macroseismic studies in the Himalaya have used the areal extent of isoseismal

radii to develop attenuation relationships (Ambraseys and Douglas, 2004). Figure 3.3 shows a comparison of

these results with the attenuation relationship developed here for the Himalaya.

While the attenuation relationship developed in this study disagreeswith that derived byAmbraseys

and Douglas (2004) at the 2σ level, the two attenuation relationships are not grossly inconsistent for inten-

sities greater than IV. Both relationships appear to parallel each other before diverging below intensity

III. I consider the sharp divergence between these relationships below intensity III to be caused by differ-

ences in the definition of the radius of perceptibility between the EMS-98 scale and the MSK scale. In fact,

28

Table 3.2: Intensity attenuation relationship coefficients obtained by other investigations used in this pa-per. Columns a, b, c, and d refer to the variables in equation (3.1). The form of the attenuation relationshipused by Atkinson and Wald (2007) and its associated coefficients are listed in Table (1) and equation (1) inAtkinson and Wald (2007). (a) This parameter was defined to be zero.

NumberArticle of a b c d

EventsBakun and Wentworth (1997) 22 3.67 1.17 0(a) -3.19

Bakun et al. (2003) 28 1.41 1.68 -0.00345 -2.08Ambraseys and Douglas (2004) 23 0.46 1.54 -0.004 -2.54

1

2

3

4

5

6

7

8

Inte

nsity

0 200 400 600 800 1000

Distance (km)

this study

Ambraseys and Douglas (2004)

Figure 3.3: Intensity attenuation with distance for a hypothetical M 6.5 Himalayan earthquake from thisstudy (solid line) and from Ambraseys and Douglas (2004) (dashed line). Intensity data from this study arein EMS-98 and data from Ambraseys and Douglas (2004) are in MSK. Error bars are 2σ.

29

the two attenuation relationships can be brought into excellent agreement by either decrementing the

value of a in this study by 0.5 intensity units or decreasing the epicentral distance by 25 km. Reasons for

this shift between the relationships could include the use of half-unit intensities in Ambraseys and Douglas

(2004), a slight bias in assessed intensities between the two studies, variations in the precision of the epi-

central locations of the calibration events between the studies, and differences in the methodology used

to calculate the calibration curves.

Of these possibilities, I can only test for the presence of a bias between the two data sets. I have com-

piled a direct comparison of 95 intensities from 3 earthquakes with common locations in both the present

catalog and Ambraseys and Douglas (2004) (Figure 3.4). This comparison indicates that the two studies are in

good statistical agreement, with more than 88% of the assessed intensities differing by no more than one

intensity unit. Although none of these earthquakes are used in the generation of the calibration curves,

this comparison shows that while most assessed intensities are identical between studies (∆ Intensity =

0), there appears to be a slight bias towards lower values in the catalog by no more than one intensity

unit. Since a bias towards lower values in intensities in the catalog would require incrementing the value of

a I may rule out the possibility that a systematic bias is responsible for the discrepancy between the two

attenuation relationships.

It has generally been assumed, based on overall similarities between the crustal structure and age

of eastern North America and India, that the regions are characterized by similar attenuation of seismic

waves and intensities (Johnston, 1996; Talwani and Gangopadhyay, 2000; Ellis et al., 2001). However, previous

authors have inferred systematic differences in both peak ground motion attenuation and weak-motion

attenuation between eastern North America and other stable continental regions worldwide (Bakun and

McGarr, 2002;Miao and Langston, 2008). Both Bakun et al. (2003) and Atkinson andWald (2007) have developed

relationships between intensity and epicentral distance for eastern North America. Figure 3.5 compares

intensity attenuation relationships in India with those from eastern North America for a hypotheticalMw

= 6.5 earthquake. For all epicentral distances, the attenuation relationship of Bakun et al. (2003) predicts

higher intensity observations in eastern North American compared to cratonic India. In contrast, the

relationship developed by Atkinson and Wald (2007) agrees with that for cratonic India above intensity V,

30

−4

−3

−2

−1

0

1

2

3

4

∆ Inte

nsity

1 2 3 4 5 6 7 8 9

EMS−98 Intensity

0 25 50

Frequency

1803 Uttarkhand1819 Allah Bund1833 Nepal

Figure 3.4: Comparison of assessed intensities at 95 common locations from the catalog and Ambraseysand Douglas (2004) for 3 earthquakes. For the histogram, the x-axis (top) corresponds to the normalizedfrequency of the combined intensity differences. For individual earthquakes, x-axis (bottom) correspondsto the assessed intensity value from the catalog. The y-axis corresponds to the difference between theassessed intensities from the catalog and those from Ambraseys and Douglas (2004) with negative valuesindicating that the intensity from the catalog is lower than that listed in Ambraseys and Douglas (2004). Forclarity, intensities for the 1819 Allah Bund and 1833 Nepal earthquake have been artificially offset to theright by 0.1 and 0.2 intensity units respectively.

31

but below intensityV, these two relationships diverge sharply, with larger intensity values being predicted

to greater distances in eastern North America. This could be due to differences in gross crustal properties

between easternNorth American and cratonic India such that higher-mode surfacewaves (Lg) travelmore

efficiently in eastern North America. However, I note that Atkinson and Wald (2007) assume a different

functional form for intensity attenuation, one that includes non-linear magnitude terms.

A comparison ofmy results with the results of Bakun et al. (2003) could be complicated by uncertain-

ties associated with their results. In particular, the intensity values for calibration events used by Bakun

et al. (2003) have not been systematically reinterpreted, and may suffer from the same problems that for-

merly plagued available intensity values for India. To further investigate the difference revealed in Figure

3.5 I directly compare attenuation from earthquakes of similar magnitude in eastern North America and

cratonic India. For lowmagnitude earthquakes (M ∼ 4.5), themedian distance at which shaking of inten-

sity III and IV is felt is twice as far in eastern North America as compared with cratonic India (Figure 3.6).

These direct comparisons corroborate the result that attenuation is at least a factor of 2 lower in eastern

North America compared to cratonic India.

While both theHimalaya andCalifornia are active plate boundary zones, there is no reason to expect

good agreement between intensity attenuation in the two regions. Nonetheless it is interesting to com-

pare the results for these two regions. My results suggest that intrinsic attenuation is small (c=-0.0006 in

equation (3.1)) in the Himalayan region which is in agreement with the results of Atkinson and Wald (2007)

(their equivalent of c has a value of -0.0007), while Bakun and Wentworth (1997) developed the California

relationship using 22 calibration events under the assumption that intrinsic attenuation was negligible

(c=0 in equation (3.1)). This low intrinsic attenuation is indicative of a highly absorptive crust (high at-

tenuation, low Q) which is expected in a tectonically active region. Allowing for a vertical shift of up to

0.5 intensity units due to differences in the intensity scales utilized, Figure 3.7 illustrates remarkably good

agreement between the both the Californian and Himalayan intensity attenuation relationships.

32

1

2

3

4

5

6

7

8

9

10

Inte

nsity

0 200 400 600 800 1000 1200 1400

Distance (km)

this study

E. North America (Bakun et al. (2003))

CEUS (Atkinson and Wald (2007))

Figure 3.5: Intensity attenuation relationship between India from this study, the results of Bakun et al.(2003) for eastern North America, and the results of Atkinson and Wald (2007) for the Central Eastern US(CEUS) for a hypothetical M 6.5 earthquake. Indian intensity data are in EMS-98 while data from easternNorth America are in MMI. Error bars are 2σ.

33

1

2

3

4

5

6

7

Inte

nsity

0 250 500 750

Distance (km)

5 Sep 2000 Koyna (M 5.2)

18 Apr 2008 Mt. Carmel, IL (M 5.2)

(a)

1

2

3

4

5

6

7

Inte

nsity

0 250 500 750

Distance (km)

26 Nov 2007 Delhi (M 4.7)

29 Apr 2003 Fort Payne, AL (M 4.6)

(b)

Figure 3.6: A direct comparison between intensity observations from eastern North American and cra-tonic India. Eastern North American intensity data are from the USGS Community Internet Intensity MapProject, error bars represent standard error estimates of the sample median. a.) Direct comparison of themedian distance to which each intensity was observed for the 18 April 2008Mw 5.2 Mt. Carmel, IL earth-quake and the 5 September 2000 Mw 5.2 Koyna earthquake. For intensities III–VI, the median distanceis statistically larger for the Mt. Carmel, IL earthquake. b.) Direct comparison of the median distanceto which each intensity was observed for the 29 April 2003 Mw 4.6 Fort Payne, AL and the 26 November2007 Mw 4.7 Delhi earthquake. Although the Delhi earthquake is larger than the Fort Payne earthquake,the median distance to which intensities II–V are smaller in India. This suggests that the attenuation dif-ference between eastern North American and India is equivalent to a magnitude increase of at least 0.2Mw.

34

1

2

3

4

5

6

7

8

9

10

Inte

nsity

0 200 400 600 800 1000 1200 1400

Distance (km)

this study

California (Bakun and Wentworth (1997))

California(Atkinson and Wald (2007))

Figure 3.7: Intensity attenuation relationship between the Himalaya from this study, the results of Bakunand Wentworth (1997) for California, and the results from Atkinson and Wald (2007) for California for a hy-pothetical M 6.5 earthquake. Indian intensity data are in EMS-98 while data from California are in MMI.

35

3.4 Estimation of Historical Epicenters and Magnitudes

The precise locations of historical earthquakes in India and the Himalaya have important conse-

quences for recurrence interval studies as well as seismic hazard assessment. Using the intensity attenua-

tion relationships derived in the preceding section, I determine the locations andmagnitudes of historical

events, examine the uncertainties of epicentral locations and magnitudes, assess the completeness of the

catalog, and take a closer look at four historical earthquakes that have previously been interpreted as great

earthquakes. A list of the location andmagnitude of historical earthquakes calculated using data from the

the catalog appears in Appendix B. Finally, I use the intensity distribution for the 2001 Bhuj earthquake

to investigate what one would infer for this event, had it been known only from historical sources.

3.4.1 Epicentral Locations and Magnitudes of Historical Events

For earthquakes prior to 1890 the only information available to us for assessing the location and

magnitude of most historical earthquakes in India comes from felt intensity data. The exceptions are

for those earthquakes whose location can be constrained from independent observations such as tide

gauge data (e.g. the 1881CarNicobar earthquake (Ortiz andBilham, 2003)), documented surface rupture (e.g.

the June 1505 central Himalayan earthquake for which surface slip has been measured (D. Yule, personal

communication, 2007)), and obvious surface deformation, (e.g. the 1819 Allah Bund earthquake (Oldham,

1926) which caused local uplift and a large region of subsidence).

The catalog affords us the possibility of refining both the location and magnitude of many earth-

quakes in the historical record. Although the approach outlined in Section 3.2 offers a sophisticated

method to quantitatively evaluate a probable epicentral location and with it, a probable magnitude, I

have found that the Bakun andWentworth (1997) algorithm frequently chooses erroneous values where the

results can be compared with instrumental values. For 100 test earthquakes for which I have both inten-

sity data and an instrumental location and magnitude, the median location error is 120 km with a median

magnitude overprediction error ofMw 0.4.

The reason for the errors in location follows partly from a paucity of observations and their spatial

36

coverage, partly from the absence of a large range of intensity values in a given earthquake and partly

from the measure of dispersion chosen as my metric in equation (3.2). Even for some very well recorded

earthquakes that do not have these shortcomings, the estimated epicentral location is often counterintu-

itive, and where I can test its true location, demonstratively incorrect. Examples are discussed below. It

may be possible to decrease the discrepancies in epicentral location andmagnitude by choosing ameasure

of dispersion that is more robust than equation (3.2) in the presence of outliers.

Where azimuthal felt-intensity coverage is limited to one quadrant, or to two contiguous quad-

rants, from the epicenter, as for example, in earthquakes near the coast, or on the southern edge of the

Tibetan plateau where reporting is inevitably one sided, there is often a trade-off between magnitude

and location. I found that location accuracy in such cases can be improved by selecting the preferred

hypocentral location to coincide with the location of the minimum magnitude, M , from equation (3.2).

This minimummagnitude location rarely corresponds to the minimum deviation location determined us-

ing equation (3.2). Lest too much credibility be attached to the coordinates derived from the minimum

deviation solution, I also list coordinates for the minimum magnitude in Appendix B. The mean location

error using the minimum magnitude location as a conservative constraint more than halves the misfit

for the 100 test earthquakes to 44 km in position; however, this method also systematically underpredicts

earthquake magnitudes byMw 0.6.

As an example, I show the location errors from the minimum deviation method for aftershocks

following the 10 December 1967 Koyna earthquake and nearby earthquakes (Figure 3.8(a)). Some earth-

quakes were misplaced out to sea, or far inland, with a median mislocation error of 120 km. For some

aftershocks, magnitudes are estimated to be larger than the mainshock. In contrast, the location of the

minimum magnitude yields a median mislocation error of 26 km (Figure 3.8(b)), with magnitudes that

were within 0.35Mw of their instrumental values. For earthquakes with more than 100 felt observations,

the location error is less than or equal to the grid spacing (∼ 9 km).

While it is clearly to some extent a subjective decision whether to use the minimum magnitude

or the minimum deviation solution, I note that choosing the minimum magnitude is consistent with the

probability that had the magnitude been larger, in many cases, it would have been felt by people in other

37

73˚ 74˚ 75˚ 76˚ 77˚15˚00'

17˚30'

20˚00'

50 km

Thane

Koyna

Killari

Marathwada

(a) Minimum Deviation Method73˚ 74˚ 75˚ 76˚ 77˚

15˚00'

17˚30'

20˚00'

50 km

Thane

Koyna

Marathwada

(b) Minimum Magnitude Method

Figure 3.8: Comparison of the epicentral misfit for instrumentally recorded earthquakes in the Koynaregion of India. On both figures, the arrow points from the instrumental epicenter towards the intensityderived epicenter. (a) Epicentral misfit in the Koyna region using the location of theminimumof equation(3.2) as the epicentral estimate. (b) Epicentralmisfit in theKoyna regionusing the locationof theminimumM from equation (3.2).

38

quadrants.

In a search for a simple discriminant to reject aberrant solutions, I found that the most accurate

locations (within 30 km of the epicenter) are those for which the locations chosen by equation (3.2) and

the minimum magnitude location differ by less than 30 km. However, if one were to apply this criterion

strictly, one would reject the locations of more than 23 of the earthquakes. I prefer to include solutions

for a larger set of events, but it is important to note the uncertainties discussed above when utilizing the

solutions presented here.

3.4.2 Catalog completeness

Using themagnitudes I have calculated for 234 events in the the catalog, I compare theirmagnitude

distribution to the magnitude distribution from the ISC catalog covering the same geographic region dur-

ing the period 1980–2000 (Figure 3.9, see Acknowledgements). Two first-order observations are apparent,

that the earthquake listing in the the catalog appears to be significantly incomplete even forMw 7.5 and

there appear to be too many earthquakes withMw > 8.

To investigate the extent to which missing aftershocks from large earthquakes might be respon-

sible for the incompleteness of the catalog below Mw 7.5, I remove known aftershocks from the catalog.

Then for each earthquake, I add aftershocks according to a Gutenberg-Richter distribution (Gutenberg and

Richter, 1954), with the largest aftershock in each sequence 1.2 units smaller than its mainshock (Bath,

1965). The resulting distribution is significantly closer to the distribution inferred from the ISC catalog,

although the distribution of events still appears to be incomplete by a factor of three for Mw 7, and a

factor of five forMw 5.

The overabundance of earthquakes withMw > 8 is likely due to tendency of the minimum devia-

tion method to over predict magnitude by nearly Mw 0.4. Although some of the catalog incompleteness

above Mw 7 could be remedied by adjusting higher magnitudes downward, it is impossible to determine

precisely which historical earthquakes have inflated magnitudes. Therefore it is not possible to correct

for this bias. Still, this bias cannot account for missing earthquakes with Mw < 6. Therefore I conclude,

that a substantial number of earthquakes are missing from the historical record. This result is not sur-

39

prising given the especially scanty early historical record that is available for some of the remote parts of

my study area. Nonetheless, assuming it is reasonable to include missing aftershocks, the distribution of

magnitudes provides a basis for quantification of overall earthquake rates for seismic hazard assessment.

Recent studies have identified surface scarps that appear to have been generated by extremely large

megathrust earthquakes (e.g. Lave et al., 2005; Kumar et al., 2006). One can use the inferred magnitude-

frequency distribution to explore the expected rate of events that are larger than those in the historical

record. Using a maximum-likelihood method (Bender, 1983) to fit the ISC results for Mw ≤ 7.5 and my

results for larger events, I infer log10(N) = 7.17− 1.034M . Although at some point one expects a simple

linear extrapolation to not be valid, this equation predicts one Mw 9.5 event in the region on the order

of once every 450 years.

3.5 Case Studies

3.5.1 The 1803 Uttarakhand Himalaya Earthquake

On 1 September 1803, a large earthquake shook much of the central Himalaya and nearby Ganges

plains causing severe damage to the town of Uttarkashi (Barahat). This earthquake is famous for its al-

leged damage to the Qutab Minar in Delhi, a structure that had stood, undamaged, since its construction

in the 13th century. This earthquake is described briefly by Mallet and Mallet (1858) and Oldham (1883).

While Seeber and Armbruster (1981) consider it the first of four great, colonial Himalayan earthquakes, no

quantitative evaluation of this earthquakesmagnitude was attempted before Ambraseys and Jackson (2003),

who compiled intensity reports fromover 30 locations and assigned a tentativemagnitude ofMs 7.5. Sub-

sequent authors (Rajendran and Rajendran, 2005; Ambraseys and Douglas, 2004) also assigned magnitudes in

the mid-7s using both Frankel’s method (Frankel, 1994) and an isoseismal area method tailored to India.

Ambraseys and Douglas (2004) assign an epicentral location near the Tibetan border (31.5N, 79.0E), while

Rajendran and Rajendran (2005) assign an epicentral location near Srinagar (Sirmur) based on the region

of maximum shaking intensity. Using the methods outlined in Section 3.2 I derive a magnitude of M 7.3

with an intensity center (epicentral location) of 30.656N 78.784E (Figure 3.10). The preferred epicentral

40

-3

-2

-1

0

1

2

3

4

4 5 6 7 8 9

Log

10(N

um

ber

of eart

hquakes p

er

year)

Mw

1980-2000 ISCHistorical Data + G-R Aftershocks

Historical Data, no aftershockslog10(N) = 6.97 - 1.0Mw

Figure 3.9: Frequency-magnitude plot of earthquakes occurring on the Indian subcontinent. Filled circlesrepresent events from the ISC catalog during 1980–2000. Diamonds represent events from the catalog;open circles with synthetic aftershock sequences added. Dashed line represents a frequency-magnituderelationship with a b value of 1.0

41

location lies 9 km south of the 1991M 6.8 Uttarkashi earthquake and 65 kmwest of the 1999M 6.6 Chamoli

earthquake. This study confirms that this was not a great earthquake, despite it being reported in numer-

ous locations throughout the Ganges Plain. The surprising proximity of the 1803 and 1991 earthquakes

is suggestive that one may be a recurrence of the other. In 188 years, the present day convergence rate

would result in a slip deficit of greater than 3 m, more than sufficient to drive a M 6.8 (Jade et al., 2004).

3.5.2 The 1819 Allah Bund Earthquake

The 16 June 1819 Allah Bund earthquake is one of the earliest events with well-documented surface

faulting (Oldham, 1926) and was responsible for the formation of Lake Sindri, a 20 km N-S by 30 km E-W

basin in the northwestern Rann of Kachchh. Upon formation, the lake flooded the village of Sindri and

destroyed a fort of the same name. Simultaneously, a region with 6 km N-S width and with an inferred

E-W length of 50–80 km rose and dammed the Puram River for several years before a flood incised the

uplifted region and the river reoccupied its old channel. This raised region was named the Allah Bund

(literally, dam of God) to distinguish it from the several artificial dams across the Puram River (Oldham,

1926). Although both the amplitude and extent of surface deformation in 1819 has been questioned (Rajen-

dran and Rajendran, 2001), the sense of the observed surface uplift and subsidence provides an approximate

constraint on the mechanism and magnitude of the earthquake (M 7.7 ± 0.2), from which an epicenter

several kilometers north of the Allah Bund has been proposed (Bilham, 1998).

Less than 200 years later, the occurrence of a second large earthquake on the Kachchh mainland,

the 2001 Bhuj earthquake (Mw 7.6), provided a much denser sampling of over 350 felt reports (the catalog

and Pande and Kayal, 2003). The similarity of these reports to the felt reports of the 1819 earthquake caused

Hough et al. (2002) to conclude that the 1819 and 2001 earthquakes were of similar magnitude. In contrast,

Ambraseys and Douglas (2004) list the 1819 earthquake as being much larger (Mw 8.2), although they note

that no detailed reevaluation of the earthquake was undertaken.

When applied to the 1819 intensity data, the algorithm outlined in Section 3.2 unexpectedly iden-

tifies an epicentral location 40 km NE of the 2001 Bhuj epicenter (Figure 3.11). This location is 100 km E

of the channel incised through the Allah Bund first described by (Burnes, 1835) and close to the mapped

42

77˚ 78˚ 79˚ 80˚ 81˚

29˚

30˚

31˚

32˚

1803 1999 Chamoli

1991 Uttarkashi

Ganges R

iver50 km

Figure 3.10: The location of the 1803 Uttarkashi earthquake as determined by the method outlined inSection 3.2. The contours represent the 50% and 67% confidence contours as determined by Bakun (1999).The instrumental epicenters of the 1991 Uttarkashi and 1999 Chamoli earthquakes (stars) are shown forreference. The location of the 1803Uttarkashi earthquake as determined byAmbraseys andDouglas (2004) isillustrated by a square. I reject the alternative epicentral location permitted by the data near the Ganges(indicated by the closed 50% and 67% confidence contours). Filled circles indicate the locations of feltreports for the 1803 earthquake within 250 km of the epicenter.

43

Island Belt Fault (Figure 3.11). The intensity-based magnitude for the 1819 earthquake is thus larger and

the epicenter more to the east than those estimated from geological or geodetic interpretations adopted

in previous studies. In contrast to the Koyna aftershocks discussed earlier, the minimum magnitude so-

lution lies south of the Kachchh mainland and is considerably less probable than the epicenter chosen by

the method of outlined in Section 3.2 given the current understanding of the regional tectonics.

Table 3.3: Epicentral locations and intensity magnitudes (MI ) of the 1819 Allahbund earthquake deter-mined using the method outlined in Section 3.2. Uncertainty in descriptions of damage to the towns ofBaliari and Umarkot inMacMurdo (1823) permit a range of EMS-98 intensities with a resulting range in theepicentral location and magnitude for the 1819 earthquake (Figure 3.11).

EMS-98Latitude Longitude Depth (km) MI Baliari Umarkot23.67 70.58 15.00 8.0 5 523.77 70.56 15.00 8.0 6 623.85 70.39 15.00 8.1 7 624.12 70.21 15.00 8.2 8 7

I find however, that the optimal epicenter is sensitive to the values of intensities assigned to points

north of the epicenter. Three of these points are mentioned telegraphically byMacMurdo (1823) and lend

themselves to debate. The catalog conservatively assigns intensity V to the southern two locations based

on the statement byMacMurdo that “shaking there was less severe than on the Kachchhmainland.” How-

ever, MacMurdo did not personally travel north of the Rann of Kachchh and damage to masonry forts on

the Kachchh mainland near Anjar and Bhuj suggest intensities as high as IX (Bilham, 1998; Ambraseys and

Douglas, 2004). Thus, intensities to the north of the epicenter could reasonably be as high as VIII and still

remain consistent with MacMurdo’s assertion.

Accordingly, I experimentally examined the shift in location caused by increasing the assigned in-

tensities at the two closest location just north of the Bund (Table 3.3). The resulting shifts in epicentral

location illustrate how sensitive the solution is to the sparse northern data. In each case, the minimum

magnitude location lies in the Gulf of Kachchh and yields a magnitude of Mw 7.6. This location can be

dismissed as inconsistent both with recent microseismic and tectonic data and with available historical

information. If the intensities at Baliari and Umerkot are arbitarily increased, the preferred epicentral

44

68˚ 69˚ 70˚ 71˚ 72˚ 73˚22˚

23˚

24˚

25˚

7.6

7.8

8

8.2

8.2

8.4

8.6

Allah Bund

Island Belt Fault

2001 Bhuj

8.08.0

8.1

8.2

Baliari

Umarkot

25 km

Gulf of Kachchh

Figure 3.11: Possible locations for the 1819 Allahbund earthquake as determined by the method outlinedin Section 3.2 (open arrows with calculatedMw). The parameters of these possible locations are listed inTable 3.3. The location of the fault responsible for the 2001 BhujMw 7.6 earthquake (Schmidt and Burgmann,2006) as well as the location of the Allah Bund fault (Malik et al., 2001) are shownwith barbs on the hangingwall. The location of the inferred Island Belt Fault is shownwith a dashed line (Malik et al., 2001). Contoursrepresent magnitudes from the epicentral location algorithm (Section 3.2) using the raw intensity data;they indicate aminimummagnitude location in the Gulf of Kachchh. The locations of Umarkot and Baliariare shown for reference. Filled circles represent felt intensity locations within 300 km of the epicenterand arrows indicate the change in epicentral location due to changes outlined in Table 3.3.

45

location passes north of the easternmost projection of the Allah Bund and the magnitude increases, even-

tually attaining an estimated magnitude ofMw 8.2.

The credibility of these solutions, however, is diminished by the disquieting sensitivity of the so-

lution to intensities north of the Allah Bund and the complete absence of observations to the west. It is

doubtful that knowledge of the shaking intensity at these locations, or locations to thewest and northwest

of the Bund will be significantly improved in the future due to an absence of detailed historical records

in the deserts of Sindh and Rajasthan. Thus, although my analysis using the methods outlined in Section

3.2 favors 8.0 ≤ Mw ≤ 8.2 and a location to the east of the Allah Bund, I am skeptical of the result due to

deficiencies in the observations. Of note, however, is the result that the magnitude corresponding to the

minimum deviation location appears to overestimate the probable true magnitude.

3.5.3 The 1833 and 1866 Nepal Earthquakes

On 26 August 1833, three earthquakes shook the Kathmandu Valley, the first, sufficiently alarming

to bring people out of doors, the second, 5 hours later to keep them there, and the third andmost destruc-

tive occurring just 15 minutes later. Bilham (1995) estimated the 1833 mainshock to be M 7.7± 0.2 using

the methods of Johnston (1996), while Ambraseys and Douglas (2004) calculate a magnitude of Mw 7.6 with

an epicenter 40 km east of Kathmandu (27.7N, 85.7E). I infer a preferred magnitude of M 7.3± 0.1 with a

location nearly 80 km ESE of Kathmandu (27.553N, 85.112E) (Figure 3.12). The calculated location roughly

corresponds to the location inferred by Bilham (1995); however, the calculated magnitude is smaller than

that inferred by both Bilham (1995) and Ambraseys and Douglas (2004). Although epicenters for the two

foreshocks are poorly constrained, using the assumption that they occurred within the source region of

the main shock yields magnitudes of M 5.1 and M 6.5, respectively.

A moderate earthquake occurred on 23 May 1866 near Kathmandu that is mentioned by several

authors (Oldham, 1883; Khattri and Tyagi, 1983; Khattri, 1987; Rajendran and Rajendran, 2005). Khattri (1987)

assesses the magnitude of the 1866 event as M 7.6 based on rupture length-magnitude scaling relation-

ships (Wyss, 1979). Although the epicentral location is poorly constrained due to a lack of observations

north of Kathmandu, the data are consistent with an epicentral location within 80 km of Kathmandu and

46

83˚ 84˚ 85˚ 86˚ 87˚25˚

26˚

27˚

28˚

29˚

1833

1866

Kathmandu

Ganges River

50 km

Figure 3.12: The locations of the 1833 and 1866 Nepal earthquakes as determined using the method out-lined in Section 3.2. The contours represent the 50% and 67% confidence regions obtained using methoddescribed by Bakun (1999). The previous estimate of epicentral location for the 1833 earthquake from Am-braseys and Douglas (2004) is represented by a square. Filled circles indicate the locations of felt reports forthe 1833 and 1866 earthquakes within 250 km of Kathmandu.

47

a magnitude of 7.2± 0.2 (Figure 3.12). Thus, according to my intensity analysis, the 1833 and 1866 earth-

quakes both appear to have ruptured similar locations in the Nepal Himalaya with similar magnitudes. In

this case, unlike the 1803/1991 earthquake pair, the slip in the second event would not have developed

over the course of 33 years with a geodetic convergence rate of 18 mm/yr (Jade et al., 2004).

3.5.4 The 2001 Bhuj Earthquake (Mw 7.6)

The 26 January 2001 Bhuj, India earthquake is the largest calibration event that I used to determine

attenuation for cratonic earthquakes, and the only cratonic calibration event above magnitude 7 (Figure

3.2). Although it appears to be circular reasoning to use the inferred attenuation relation to determine

an optimal location and magnitude for this earthquake, this is a potentially interesting exercise because

the attenuation relation is primarily constrained by events with Mw ≤ 6 (Figure 3.2). The intensity

derived location for the Bhuj earthquake using intensity values from the catalog yields a location only

12 km away from the instrumental epicenter. This is slightly larger than the grid spacing (9 km) used in

the epicentral location method. However, the magnitude is estimated as Mw 8.0 using the attenuation

relationship derived using only earthquakes from the Indian craton, and Mw 8.6 using the attenuation

relationship derived for all of India. Thus, even though the Bhuj intensities are used to constrain the

attenuation relation, the method of Bakun andWentworth (1997) over-predicts the magnitude of this event

by 0.4 or 1.0Mw units.

To explore why I obtain an unreasonably large magnitude for the Bhuj earthquake (and by implica-

tion, an uncertain magnitude for the nearby 1819 Allah Bund earthquake), I examine the intensity values

for the 2001 earthquake as a function of distance. The decay in intensity with distance shows a systematic

difference betweenwith the intensities anticipated by equation (3.1) for aMw 7.6 earthquake (Figure 3.13).

Moderate to small intensity observations are found at significantly greater distances than those predicted

by the attenuation relationship and, in particular, median intensity observations between 200 km and 875

km (median distances for intensities IV–VII) appear between one-half to one intensity unit greater than

anticipated.

Four possible explanations for the discrepancy illustrated in Figure 3.13 are:

48

2

3

4

5

6

7

8

9

10

EM

S−

98

0 150 300 450 600 750 900 1050 1200 1350 1500

Distance (km)

2001 Bhuj Mw 7.6

0

25

50

Fre

quency (

%)

12345678910

MSK

0

25

50

0 500 1000 1500

Distance (km)

Figure 3.13: Intensity observations of the 2001 BhujMw 7.6 earthquake compared to the attenuation curvederived for cratonic India for an earthquake of Mw 7.6. Open circles represent observed intensities, dia-monds represent the median distance for each observed intensity level. Dashed lines represent the 2-σenvelope of uncertainty in the intensity attenuation model as a function of distance.

49

(1) That intensities for the Bhuj earthquake were systematically over-estimated.

(2) That there is, or can be, a non-linear dependence of intensities on magnitude for large earth-

quakes.

(3) That intensity observations at regional distances are amplified by the presence of higher-mode

surface (Lg) waves.

(4) That the intensity observations for the Bhuj earthquake indicate a high-Q in the Kachchh Basin.

I shall address each possibility in turn. First, I consider it unlikely that the intensities for the Bhuj

earthquake were systematically over-estimated. Most of the values for Bhuj in the catalog are, in fact,

systematically lower than the values inferred by Hough et al. (2002) whose intensity assignments were

based onmedia reports andhave been shown to be biased relative to those estimated fromdirect surveying

of damage (Hough and Pande, 2007).

Second, the functional form of equation (3.1) is identical to the functional form assumed for atten-

uation of peak ground acceleration (Evernden et al., 1973). Several studies have shown a good correspon-

dence between intensity and instrumentally determined ground motion measures (e.g.Wald et al., 1999b).

One might therefore reasonably expect equation (3.1) to be appropriate for characterizing intensity at-

tenuation for large events. Short of significant non-linearity associated with ground motions at sediment

sites, equation (3.1) appears to be appropriate for characterization of peak ground acceleration for large

and small earthquakes.

Third, when higher-mode surface wave trains develop and propagate in the continental crust, the

highest amplitude shaking typically has a long duration. It is thus reasonable, if not expected, that a pro-

longedLg wavetrain with a given peak acceleration will produce a higher perceived intensity observation

than groundmotions with the same peak acceleration and amuch shorter duration. Shaking durationwill

clearly be a potential factor for structural damage; it is self-evident that marginally perceptible shaking

is more likely to be felt if the strongest amplitudes are prolonged. Clearly, human perception of higher-

mode surface waves decreases with distance from the epicenter by a non-integer amount. Thus, a simple,

uniform adjustment of intensity observations to correct for amplification is not possible.

50

Lastly, I can consider the possibility that the intensity distribution reflects especially high-Q in the

Kachchh Basin. Bodin et al. (2004) calculate Q for the Kachchh Basin using aftershocks of the 2001 Bhuj

earthquake and note that their estimates are higher than estimates of Q in northern India calculated by

Singh et al. (1999). In contrast, in a regional study of Lg attenuation, Pasyanos et al. (2009) shows values

of Q in the Kachchh Basin to be closer to those measured in northern India (Singh et al., 1999). Addition-

ally, results in Mitra et al. (2006) indicate that estimates of Q from the 2001 Bhuj earthquake itself are

systematically higher than estimates from other regional events. To test the possibility that the attenua-

tion properties of the Kachchh Basin affect intensity observations and consequently inflate the calculated

magnitude of the 2001 Bhuj earthquake, I removed all intensity observations within 200 km of the Bhuj

epicenter and inverted for magnitude. The removal of all observations within 200 km of the epicenter

results in an increase in epicentral location uncertainty, but essentially no change in magnitude. This is

not surprising since equation (3.1) indicates that a change in magnitude will have a larger influence on

distant, lower intensity observations than near-source high intensity observations.

For large earthquakes, locations are well determinedwhere sufficient spatial coverage exists. How-

ever, the magnitudes of large events are not well determined using the method outlined in Section 3.2,

and requires consideration. As shown in Figure 3.13, the intensity values do not generally match the

intensities predicted by equation (3.1) for a Mw 7.6 earthquake between distances 200 km and 875 km

from the epicenter. In addition, the inferred intensity distribution includes moderate intensities to sig-

nificantly greater distances than the predicted distribution. The lowest felt intensities (II–III) similarly

extend farther than predicted. These results suggest that the formation of higher-mode surfaces waves

due to long shaking durations in a high-Q environment have amplified intensity observations at regional

distances. Simple correction of this amplification is not possible and these results moreover provide a

caution regarding the use of the Bakun and Wentworth (1997) method with an attenuation relation of the

form given by equation (3.1). In particular, if the attenuation relation is constrained largely or entirely by

small or moderate earthquakes, the magnitudes estimated for large historical earthquakes can be grossly

overestimated.

51

3.6 Discussion

The determination of the magnitudes of historical earthquakes is of interest since, were a com-

plete inventory of historical earthquakes available, I could subject a region to investigations of moment

release over space and time. Statistical tests assuming a Gutenberg-Richter distribution of magnitudes

show that I am missing 30% of the moderate earthquakes during the period for which most of data are

derived (1800–2000). Thus, while moment-release studies can be undertaken for the entire region, they

are doomed to be less reliable on a local scale, in particular for the relatively frequent 6.5 ≤ Mw ≤ 7.5

events that are typically important for controlling probabilistic hazard.

The caveat discussed in the previous section, that attenuation findings for small earthquakes do not

provide satisfactory predictions for the attenuation observed in the largest earthquakes and therefore

unsatisfactory magnitude predictions, is perceived to be a substantial problem in India, since it is for

these largest earthquakes that reliable magnitude information is most needed. It is possible that a similar

problem exists with the intensity relations established for North America.

Epicentral locations determined using the methods I describe show demonstrable scatter. One

question that arises in the determination of magnitude and epicentral location for early earthquakes is

what constitutes an acceptable determination of these parameters. If I am interested in establishing an

inventory of potentially active faults, I should presumably prefer locations that lie within one source di-

mension, e.g. a fault rupture length, of the earthquake. If I am interested in identifying segments of those

faults that remain unruptured, I require yet higher accuracy. It is clear from the analysis I present here

that few of the post-1950 solutions for earthquakes with Mw < 6 are within one fault length of the in-

strumental epicenter, and by implication I must assume that the same is true of the earthquakes earlier

in the catalog.

For many of the events in the catalog with poor intensity coverage, I do not attempt to determine

a location using quantitative methods. Yet even for these earthquakes I recognize that the approximate

location and its intensity data are of utility in seismic hazard studies. Chapter 2 utilizes this information

to map maximum shaking intensity encountered in a grid throughout India.

52

For larger earthquakes (7 < Mw < 8) where rupture lengths range from 30 km to 300 km I find, in

general, that the preferred location lies near or above the inferred rupture surface, but I note that even

for some very large earthquakes in the catalog the dimensions and location of the rupture zone remain

enigmatic (e.g Chittagong 1762 and Bihar-Nepal 1934).

For 100 earthquakes in the instrumental period (post 1950) for which I have both epicentral pa-

rameters and intensity data, I find that fewer than 30% of these earthquakes can be located to within one

fault length of the true epicenter using intensity data. The median mislocation location error using the

method of Bakun andWentworth (1997) exceeds 100 km. However, choosing the minimummagnitude loca-

tion instead of the minimum deviation location reduces the misfit by a factor of two. The reason for the

poor performance of the algorithm is partly due to the small number of observations available formany of

these earthquakes as well as the small dynamic range of the intensity observations for each earthquake. I

conclude that the algorithm cannot be expected to do better for historical earthquakes; locations accura-

cies are likely to be no better than 50 km. One disappointing result is that from the data alone, there seems

to be no reliable way to characterize the quality of each solution. In general, earthquakes with fewer than

10 locations gave consistently unreliable locations. I found that the least reliable solutions were those

where large differences were found between the minimum deviation location and the minimum magni-

tude location. The best locations were found to be those in which these two locations agreed to within 30

km, but this applies to fewer than 30% of the data.

3.7 Conclusions

Newly available intensity observations for India provide a wealth of material for evaluating the

location and magnitude of numerous earthquakes that have hitherto been amenable only to qualitative

analysis, and, in particular, permit us to assess attenuation throughout the subcontinent. I use attenua-

tion relation derived from modern (post 1950) earthquakes with well-determined instrumental locations

and the method of Bakun and Wentworth (1997) to estimate the optimal locations and magnitudes for 181

historical earthquakes, with case studies of large events in 1803, 1819, 1833, and 1866.

Of particular interest are the characteristic attenuation versus distance parameters for India. I

53

quantify attenuation for all Indian earthquakes, and separately for plate boundary events (Himalaya) and

cratonic events. I find that intensity attenuation in theHimalaya region is comparable to that in California,

while attenuation in cratonic India is significantly higher than attenuation in the central/eastern United

States.

One unexpected finding is that for the largest of the cratonic earthquakes (Bhuj 2001 and Allah

Bund 1819) my attenuation relation significantly overestimates the magnitudes estimated from instru-

mental and/or geological constraints. This results from shaking being felt more strongly out to greater

distances than expected bymy attenuation relationship. I suggest that thismay be a systematic effect that

is common to all attenuation models. The distant shaking from large earthquakes is simply not well char-

acterized by shaking from small earthquakes. I propose that the duration ofLg shaking at large distances

may be responsible for this effect.

My search for uncertainty criteria to describe location accuracy is unsatisfactory in that I have

found no objectivemethod, from the intensity data alone, to quantify the accuracy ofmy solutions. Where

more than 100 intensity values are available the solution is usually within 30 km of the true epicenter.

This criteria applies to only 16 events, less than 3% of the catalog. Where the minimum deviation and

minimum-magnitude solution are close, the calculated epicenter is usuallywithin 30 kmof the instrumen-

tal location, but even this condition applies to less than 30% of the instrumental catalog, and by extension,

to fewer than 177 of the 570 earthquakes in the entire catalog.

Themagnitudes of earthquakes in the instrumental periodwere notwell characterized by the Bakun

and Wentworth (1997) algorithm. This is perhaps not too surprising since the assigned magnitude for a

given attenuation depends on distance, which as summarized above shows a large range of mislocation

errors. Magnitudes were in general overestimated by a median mismatch of Mw 0.4 for the 100 earth-

quakes for which instrumental magnitudes were known. The median magnitude misfit using the mini-

mum magnitude location underpredicts the instrumental magnitude byMw 0.6. Again, this uncertainty

in magnitude suggests that historical earthquakes cannot be characterized to better thanMw ± 0.5 from

the historical data analyzed here.

Chapter 4

Interseismic Strain Accumulation along the Western Boundary of the IndianSubcontinent

4.1 Introduction

The Chaman Fault System is a major continental plate boundary transform zone separating the

Indian and Eurasian plates. South of the Afghanistan/Pakistan border, this fault system is composed of

three major left-lateral faults, the Chaman, the Ghazaband and the Ornach-Nal faults (Figure 4.1). The

Chaman Fault proper comprises themajority (860 km) of this 1250 km long transform zonewhich connects

the Makran subduction zone to the currently inactive right-lateral Herat Fault in northern Afghanistan. I

utilize two campaign GPS transects bounded on their East by continuous GPS stations located on the stable

Indian Plate to estimate the interseismic deformation rate along the Chaman Fault System. North of the

town of Chaman, Pakistan, I utilize nearly 4.5 years of InSAR data to produce a fault-centered transect of

interseismic deformation rate. In all three locations I examined, I find shallow locking depths. In regions

with shallow locking depths, large strike-slip earthquakes (Mw > 7) are unlikely due to the reduced

seismogenic thickness. However, large earthquakes have occurred in the past century in the ranges east

of the Chaman Fault both east and south of Quetta. The absence of GPS data south of the town of Chaman,

Pakistan hinders the ability to estimate interseismic deformation rates in this region.

4.2 Tectonic Summary

The Ornach-Nal fault is the southernmost on-land segment of the Chaman Fault System (Figure 4.1;

Zaigham (1991); Lawrence et al. (1992)). Running nearly 250 km north from the Makran Coast, the Ornach-

55

65˚

70˚

31˚

Orn

ach

Na

l F

au

lt

Ghazaband F

aultC

ham

an F

ault

Herat Fault

Quetta

Chaman

Pishin

Nushki

Sukkur

Karachi

Ab−e−Istada

Qalat

Kabul

Sulaiman Lobe

Kirthar Range

Figure 4.1: Map of the western boundary of the Indian Plate, highlighting the major faults of the ChamanFault System, place names mentioned in the text are also indicated. The map projection is oblique Mer-cator about the pole of relative motion between the Indian and Eurasian plates. Thrust faults are shownwith filled triangles on the hanging wall, all other faults shown are strike slip.

56

Nal Fault offsets primarily Cenozoicmudstones and shales (Snead, 1964; Zaigham, 1991; Lawrence et al., 1992).

The lack of detailed fieldmapping in this region, combinedwith themonotonyof themudstones and shales

have made estimates of the total offset across the Ornach-Nal fault impossible.

At the northern terminus of theOrnach-Nal fault, deformation likely stepswestward onto theGhaz-

aband Fault (Figure 4.1). Although no modern geodetic measurements exists across this segment of the

Chaman Fault System, the Ghazaband fault is a likely candidate for the source of the 1935 Quetta earth-

quake (∼ M 7.5) which caused surface cracking and extensive damage in the populated valleys south of

Quetta and approximately 15 km east of the Ghazaband Fault (Ramanathan and Mukherji, 1938; Engdahl and

Villasenor, 2002). The Ghazaband Fault parallels the Chaman fault northward to the town of Pishin where

it disappears into the southern end of the seismically quiet Katawaz Block (Haq and Davis, 1997).

South of the village of Nushki, Pakistan, the last of the arcuate thrust faults of the Makran merge

to form the N-S trending Chaman Fault (Figure 4.1). The Chaman Fault and Ghazaband Fault parallel each

other northwards to the latitude of the town of Chaman, Pakistan. Immediately north of the town of

Chaman, the fault veers NNE, enters Afghanistan and becomes the dominant fault in the Chaman Fault

System.

Lawrence et al. (1992) estimate the onset of faulting along the Chaman Fault System as Oligocene-

Miocene based on post-depositional offsets of the Eocene-Oligocene aged Khojak flysch. These displace-

ments suggest an average slip-rate of 19–24 mm/yr over the last 20–25 My. Similarly, correlation of

Pliocene volcanic units that straddle the fault north of Ab-e-Istada, Afghanistan, yields an average slip-rate

of 25–35 mm/yr over the last 2 My (Beun et al., 1979).

Kinematic and analogue modeling of the western boundary of the Indian plate indicates that pure

strike-slip motion is required along the Chaman fault with shortening being accommodated in the Su-

laiman Lobe and Kirthar ranges to the east of the Chaman fault (Haq and Davis, 1997; Bernard et al., 2000).

Estimates of modern plate motion indicate convergence of the Indian plate relative to Eurasia at a rate of

27 mm/yr at N7E near Sukkur, Pakistan (Altamimi et al., 2007). I estimate the amount of sinistral slip along

the Chaman Fault by noting that its strike between Chaman and Ab-e-Istada is approximately N34E, sug-

gesting approximately 24mm/yr of sinistralmotion and 12mm/yr of shortening are being accommodated

57

across the Chaman Fault System.

At least four strike-slip earthquakes with M > 6 have been recorded historically on the Chaman

Fault. In 1505, a strike-slip event along the Chaman fault occurred west of Kabul (Oldham, 1883; Babur,

1912; Lawrence et al., 1992; Ambraseys and Bilham, 2003b), in 1892, anM 6.5 strike-slip event occurred near

the city of Chaman (Griesbach, 1893; Ambraseys and Bilham, 2003b), in anMS 6.7 occurredmid-way between

Chaman andNushki (Lawrence and Yeats, 1979; Engdahl andVillasenor, 2002) and in 1978 anMw 6.1 occurred

north of Nushki, Paksitan (Yeats et al., 1979; Engdahl and Villasenor, 2002). Recent seismicity along the fault

appears to consist mostly of small, M 3–5, earthquakes primarily located in regions of major historical

seismicity. However, due to the lack of seismic instrumentation in the region, most earthquakes occurring

in the Chaman Fault System are poorly located.

Earthquakes along the Chaman Fault appear to consistently rupture to the surface. The 1505 earth-

quake near Kabul was noted to have surface rupture (Oldham, 1883; Babur, 1912), while surface rupture

from the 1892 Chaman, Pakistan, earthquake offset railroad tracks crossing the fault by 0.75 m (photo-

graph and diagram in Griesbach, 1893). Surface rupture along the Chaman Fault from the 1892 earthquake

was estimated to extend for at least 60 km with its northernmost mapped extent lying near the town of

Chaman, Pakistan (Griesbach, 1893; Landor, 1902). Due to its remote location, it is unknown whether the

1975 MS 6.7 Chaman Fault earthquake between Nushki, Pakistan, and Chaman, Pakistan, produced sig-

nificant surface rupture. Finally, field investigations by Yeats et al. (1979) indicate that the 16 Mar. 1978

Mw 6.1 near Nushki, Pakistan, also ruptured to the surface.

Based on analysis of seismicmoment release along the Chaman fault system over the last 150 years,

both Bernard et al. (2000) and Ambraseys and Bilham (2003b) argue that a significant slip deficit exists along

the Chaman fault, particularly north of∼ 31◦ latitude.

58

4.3 Methods

4.3.1 GPS

GPS measurements in Pakistan are historically of limited coverage and duration with no measure-

ments prior to 2001. Campaignmeasurements from 13 sites with locations predominantly north of Quetta

as well as 8 sites along the Makran coast and inland to Panjgur, Pakistan, have been measured at least

twice in the period 2006–2010 and compared to continuousmeasurementsmade in Karachi (KCHI), Sukkur

(SIBA), Peshawar (NCEG) and Quetta (QTAG and QTIT) (Figure 4.2).

The location of campaign GPS measurements across both the Chaman and Ornach-Nal faults are

constrained by the location of major roads to provide security and ease of access. The continuous points

in Pakistan are operated from flat-roofed concrete frame buildings and the campaign points aremeasured

on bipods set on stainless steel screws cemented into exposed rock. GPS observations were recorded with

Trimble NetRS, 5700 and R7 receivers using a 30 s sampling rate, and processed using an elevation cutoff

angle of 10◦. Campaign data have durations of 3–7 days from each site.

Daily data from all sites were processed along with data from at least 4 continuous stations in Pak-

istan and 10 regional IGS stations using GAMIT version 10.35 (King and Bock, 2002). These regional solutions

were then combined with global solutions from SOPAC (http://sopac.ucsd.edu) using GLOBK/GLORG

version 5.17 (Herring, 2002) to determine time series and velocities consistent with the ITRF2005 reference

frame. To achievemore realistic uncertainties, velocities were re-estimated using randomwalk noise esti-

mates obtained from iterative time series modeling. Final velocities were then transformed into an Indian

plate-fixed reference frame using pole-of-rotation parameters published in Altamimi et al. (2007).

4.3.2 InSAR

I acquired 27 ascending pass SAR scenes spanning nearly 4.5 years from track 213 frame 621 of the

European Space Agency’s Envisat satellite (Figure 4.2A, 4.3). A 90 meter resolution DEM was constructed

from SRTM version 2 data (Farr et al., 2007) and used to remove topographic fringes. Interferograms

were produced using the ROI PAC InSAR software package developed at the Jet Propulsion Laboratory

59

62˚ 64˚ 66˚ 68˚ 70˚ 72˚

24˚

26˚

28˚

30˚

32˚

34˚

36˚

NCEG

TURT

KBUL

SIBA

KCHI

PANG

A

B

C

62˚ 63˚ 64˚ 65˚ 66˚

25˚

26˚

27˚

ZHAOBEDI

ORMA

PANG

PASNGWAD

SHFD LAKC

C

66˚ 67˚ 68˚ 69˚

30˚

31˚

LORI

QLAS

SANJ

CHTR

HRNIHRNI

SHRGSHRG

ZART

MUSB

KACH

SURB

QTIT

LAKP

SARN

QILA

SHBGCHMC

KHST

QTAG

B

Figure 4.2: Map showing GPS station locations and names along the western boundary of the Indian Plate.A.) GPS stations throughout Pakistan. Dashed rectangle indicates the ground footprint of Envisat track213 frame 621. B.) Stations in the Quetta Syntaxis where there is a high station density. C.) Stations alongthe Makran Coast.

60

in Pasadena, CA (Rosen et al., 2004). Interferograms were sampled with 8 looks in range and 40 looks in

azimuth to produce 160 m× 160 m resolution cells, filtered using a power spectral method (Goldstein and

Werner, 1998) and unwrapped using a least squares methodology.

From the set of 27 ascending pass scenes, I select 12 SAR image pairs that minimize both temporal

and spatial baselines aswell asminimizing atmospheric noise (Figure 4.3). These 12 scenes are then used to

form a rate-map using the iterative methodology outlined in Biggs et al. (2007). Formation of the rate-map

utilizes a pixel-based weighted least squares approach with a variance-covariance matrix that accounts

for estimates of atmospheric phase delay and orbital uncertainty from each interferogram.

To estimate variance-covariance parameters for errors due to atmospheric phase delay, I analyze

a 48× 48 km region from the tectonically quiet, southeastern corner of each interferogram. I determine

a best-fit 1-D covariance function of the form,

c(r) = σ2atme(−αr), (4.1)

where σ2atm is the variance, r is the distance between pixels and α is the e-folding length. Parameters

in Equation (4.1) are estimated from the radially averaged 2-D autocovariance function calculated using

the cosine Fourier transform of the power spectrum (Hanssen, 2001; Parsons et al., 2006). The estimated

atmospheric variance (σ2atm) and e-folding length (α) are then used toweight pixels in each interferogram

during formation of the rate-map.

Although I use ROI PAC to remove the effects of baseline separation by approximating the earth as a

smooth ellipsoid, imperfect knowledge of the satellite’s orbit occasionally results in apparent deformation

in the form of a residual tilt spanning the InSAR scene. Since the interseismic deformation signal sought is

inherently long-wavelength, estimating the residual orbital correction by simply removing a best-fitting

plane of the form

z = ux+ vy + w, (4.2)

where x and y are, respectively, the across-track and along-track directions in radar coordinates and

u, v and w are the unknowns to be estimated, would result in the removal of a portion of the tectonic

signal. To minimize this effect, I estimate orbital corrections using Equation (4.2) from regions as far

61

0

200

400

600

800

1000

Perp

endic

ula

r B

aselin

e (

m)

2004 2005 2006 2007 2008

Date

Envisat Track 213 Frame 62130 m Median Perpendicular Baseline

Mw

5 e

arth

qu

ake

( ) 1367 m

Figure 4.3: Date versus perpendicular baseline plot for Envisat track 213, frame 621. Filled circles repre-sent individual SAR scenes and solid lines represent interferograms. There is one perpendicular baselineoutlier indicated on the correct date in parenthesis along side the associated perpendicular baseline value.The vertical dashed line corresponds to anMw 5.0 earthquake on 21 Oct. 2005 along the Chaman fault inthe northern portion of Envisat track 213 frame 621. The 12 interferograms shown have amedian perpen-dicular baseline of 30 m, corresponding to an altitude of ambiguity of more than 450 m.

62

from the expected deformation signal as possible by masking out a swath 50 km wide, centered on the

fault. I then calculate variance-covariance parameters for uncertainty in the orbital correction using the

standard deviation of the parameters estimated in Equation (4.2).

4.4 Results

4.4.1 Ornach-Nal

I project the GPS velocity field of stations spanning the Ornach-Nal Fault into fault normal and fault

perpendicular directions based on the azimuth of the surface trace of faulting west of Las Bela (Figures 4.4

and 4.5). Two GPS stations (LAKC and SHFD, see Figure 4.2) straddle the subaerially exposed mud ridges

that comprise what appears to be the active trace of the Ornach-Nal Fault. However, Figure 4.5 shows

that themajority of the sinistral motion across the plate boundary occurs west of this mud ridge, between

stations SHFD and ZHAO (Figure 4.2C).

62˚ 64˚ 66˚24˚

26˚

20±2 mm/yr

Orn

ach N

al

Fault

?

Figure 4.4: GPS velocities of stations from the Makran region of Pakistan. All velocities are relative to thestable Indian Plate as defined inAltamimi et al. (2007) and are plotted using aMercator projection. The exactlocation of the offshore intersection of the subduction zone and the Chaman Fault System is unknown andis denoted with a question mark. Station names appear on Figure 4.2C

Assuming that interseismic deformation across the plate boundary is accommodated by a single

63

−20

−10

0

Velo

city (

mm

/yr)

−200−175−150−125−100 −75 −50 −25 0 25 50 75 100

Ornach Nal Fault Normal Distance (km)

Slip rate : 14.7 mm/yrLocking Depth: 7.3 km

Orn

ach−N

al F

ault

Nom

inal

Pla

te B

oundar

y

95% HPD fault zone location

PANG

BEDI ZHAO

SHFDLAKC

KCHI

Figure 4.5: GPS profile across the Ornach-Nal Fault. Velocities and uncertainties are projected into a di-rection parallel to the Ornach-Nal Fault and are relative to the stable Indian Plate. Uncertainties shownare 2σ. The thick horizontal bar indicates the 95% HPD range for possible fault locations. The dottedline represents the model that maximizes the empirical posterior likelihood function as determined us-ing a Markov-Chain Monte Carlo method (Mosegaard and Tarantola, 1995). The slip rate and locking depthfor the fault location that satisfies both the posterior likelihood and geological critera (the nominal plateboundary) are indicated on the figure.

64

fault with an unknown location, Markov-Chain Monte-Carlo (Mosegaard and Tarantola, 1995) analysis of

the velocity profile indicates that the data are most consistent with an interseismic deformation rate of

14.7mm/yr with a 95% Baysian High Posterior Density (95%HPD) region of 12.8–18.2mm/yr and a locking

depth of 7.4 kmwith awide 95%HPD of 1 km to 17.7 km. Similarly, estimation of the fault location suggests

that the plate boundary lies approximately 27 km west of the mud ridge originally thought to represent

the plate boundary (Figure 4.6). This fault location is associated with a 95%HPD region spanning nearly 35

km, indicating a large uncertainty. The majority of these locations lie beneath the large Hingjal synform

and are consistent with lineations extending north and south of the synform (Figure 4.6).

Previous authors have suggested that the mud ridges west of Las Bela are eruptive features (Jones,

1961; Bannert et al., 1992). Reanalysis of outcrop lithology by Delisle et al. (2002) as part of a systematic study

on mud volcanism in Pakistan indicates that the mud ridge (Figure 4.6) is actually an outcrop of Parkini

Mudstone, the source material erupted from mud volcanoes farther south and west of Las Bela, and not

part of an eruptive feature.

The velocity of the site at Panjgur, Pakistan, is nearly 82% of the expected ITRF2005 rate (Altamimi

et al., 2007). This observation combined with the low velocities of both SHFD and LAKC relative to the

stable Indian Plate suggests that the remaining plate boundary deformation occurs across faults on the

the Eurasian side of the plate boundary. It is also notable that the Ornach-Nal parallel velocity of the site

in Panjgur (PANG, Figures 4.2 and 4.5) is 7 mm/yr faster than the expected far-field velocity due solely to

slip on the Ornach-Nal fault. This disagreement is likely due to unmodeled convergence along the arcuate

faults that comprise the subaerial Makran forearc between Bedi and Panjgur, Pakistan. Examination of

fault normal motion indicates that there is little convergence across the Ornach-Nal Fault (< 2 mm/yr)

(Figure 4.4).

4.4.2 Chaman Fault near Chaman

I project the GPS velocity field of stations north of Quetta, Pakistan, onto a Chaman Fault perpen-

dicular profile and calculate both fault-parallel and fault-normal velocities (Figure 4.7 and 4.8). Due to

the location of the profile near the intersection of the northern Kirthar Range and the Sulaiman Lobe, the

65

SHFD LAKC

ZHAO

10 mm/yr

Bela

Jhal Jhao

Hin

glaj

Syn

form ][

?

?

?

Mud Ridge

65˚30' 66˚00' 66˚30'

26˚00'

26˚30'

25 km

Figure 4.6: Landsat 7 image of the southern Ornach Nal fault and adjacent Hinglaj synform. The largesquare brackets indicate the spatial region encompassed by the 95% HPD region shown in Figure 4.5. Geo-logically likely locations for the plate bounding fault(s) are indicated by the NE-SW trending dashed lines.The preferred plate bounding fault is the easternmost left stepping pair of faults across the Hinglaj syn-form. The gap between the fault tips corresponds to the deepest part of the synform (Bannert et al., 1992)and is likely a pull apart feature. GPS velocities are relative to the stable Indian Plate and are identical tothose shown in Figure 4.4. The image is a combination of bands 7, 4 and 2 to highlight geological informa-tion.

66

velocity field is complex and includes deformation across multiple structures. On Figure 4.8, one step in

fault parallel motion is evident between stations CHMC and SHBG. These stations straddle the Chaman

Fault and their close spatial proximity and velocity difference are consistent with 7.5 mm/yr of inter-

seismic deformation on the Chaman Fault (95% HPD range of 5.6–9.6 mm/yr) with an extremely shallow

locking depth (2.7 km with a 95% HPD range of 0–6.6 km).

66˚ 68˚

30˚

20±2 mm/yr

Ghazaband F

ault

Cham

an F

ault

Figure 4.7: GPS velocities of stations in the region of Quetta, Pakistan. All velocities are relative to thestable Indian Plate as defined in Altamimi et al. (2007) and are plotted using a Mercator projection. Stationnames appear on Figure 4.2B.

Although located only 8 km from the trace of the Chaman Fault, station CHMC is moving at approx-

imately 75% of the expected ITRF2005 velocity. This observation suggests that most of the plate boundary

motion is accommodated by faults on the Indian Plate east of the Chaman Fault. This observation is in con-

trast to observations further south of diffuse deformation being accommodated on the Eursian Plate near

the latitude of Panjgur, Pakistan. Further, the location of Sukkur (SIBA) on the stable Indian Plate com-

bined with the velocity profile in Figure 4.8 indicates that 12 mm/yr of sinistral motion is accommodated

across structures east of the Chaman Fault.

Convergence in the Kirthar range is 5.0 ± 1.3 mm/yr, as calculated using a weighted mean of the

following stations CHMC, SHBG, QILA, SURB, LAKP (see Chapter 5). Station SARN appears to be contam-

67

−20

−10

0

Ve

locity (

mm

/yr)

0 25 50 75

Chaman Fault Normal Distance (km)

Slip rate : 7.53 mm/yrLocking Depth: 2.68 km

Cham

an F

ault

Ghaz

aban

d F

ault

CHMC

QILA SARN

LAKP

SURB

KACH

SHBG

Figure 4.8: GPS profile across the Chaman Fault. Velocities and uncertainties are projected into a directionparallel to the Chaman Fault and are relative to the stable Indian Plate. Uncertainties shown are 2σ. Thedotted line represents themodel thatmaximizes the empirical posterior likelihood function as determinedusing aMarkov-ChainMonte Carlomethod (Mosegaard and Tarantola, 1995). The slip rate and locking depthfor this model are indicated on the figure.

68

inated by a large seasonal signal in its east component yielding a convergence estimate of 0 mm/yr and

thus has been disregarded. The nearly uniform convergence signal observed at stations west of Quetta

suggests that shortening in the Kirthar Range is focussed along structures east of Quetta, such as the

Dezghat-Bannh Fault System which ruptured during the 1931 Mach earthquake (Chapter 5).

4.4.3 Chaman Fault near Qalat

North of Chaman, Pakistan, the Chaman Fault enters a restraining bend and the fault trend be-

comes more perpendicular to the azimuth of ascending Envisat satellite passes. This favorable geometry

increases the amount of fault parallel motion visible in the radar line-of-sight compared with locations

further south along the Chaman Fault. I have produced a best-fitting ratemap using 12 Envisat scenes (Fig-

ure 4.9) and calculated line-of-sight displacements through binning of observations by distance from the

fault trace. Estimates of line-of-sight displacement for each bin are then calculated using a least-squares

approach (Figure 4.10).

Slip-rate and fault lockingdepthuncertainty are estimated fromFigure 4.10 using the two-dimensional

strike-slip model of Savage and Burford (1973) and a Monte Carlo resampling technique (Wright et al., 2001).

The slip-rate and locking depth estimates of 1.91 ± 0.31 rad/yr in the radar line-of-sight are consistent

with 16.8 ± 2.7 mm/yr of fault parallel motion beneath a locking depth of 5.4 ± 2.4 km. One obvious

feature seen in both the best-fit rate-map and the fault-centered profile is the deformation in the Tarnak

Rud valley near the village of Qalat. Analysis of the deformation around the Qalat area using a short-

baseline methodology shows nearly 15 mm/yr of deformation in the line-of-sight of the radar during the

time period 10 Aug. 2004–6 Jan. 2009 (Figure 4.11;Berardino et al. (2002); Hooper (2008)). Comparisons be-

tween Landsat 7 band combination 4,3,2 and the InSAR data show a good correspondence between areas of

subsidence andmodern river channels. The region of greatest subsidence is located downstreamof the re-

gion of densest agricultural usage but close to the region of densest population (Figure 4.11B), suggesting

that the most probable cause of the deformation is groundwater withdrawal. As Landsat 7 data acquired

after 31 May 2003 contains data gaps due to an instrument failure, only data prior to the short-baseline

time period are available. It is possible that in the time between when the Landsat 7 shown in Figure 4.11

69

66˚ 67˚ 68˚

31˚

32˚

−1

0

1

2

3

radia

ns/y

r

CHMC

SHBG25 km

Figure 4.9: InSAR rate-map derived from stacking 12 ascending pass Envisat interferograms. Solid arrowindicates the flight direction of the satellite while the transparent arrow indicated the line-of-sight direc-tion. Values are phase velocity in rad/yr in the line-of-sight of the radar and referenced to a pixel in thefar NW corner of the scene. More positive values of phase velocity indicate increasing radar line-of-sightdistance. Interferograms used in construction of the rate-map are indicated by solid lines in Figure 4.3and have a median perpendicular baseline of 30 m. The surface trace of the Chaman Fault is indicated bythe dashed line. For reference, the locations of GPS stations CHMC and SHBG are indicated in the south-ern portion of the map. The increasing radar line-of-sight velocities near the town of Qalat, Afghanistan,(black triangle) are likely tied to subsidence due to groundwater withdrawal for agriculture.

70

−2

−1

0

1

2

Lin

e−

of−

Sig

ht V

elo

city (

radia

ns/y

r)

−30 −20 −10 0 10 20 30

Distance from Chaman Fault (km)

1000

2000

3000

4000

Ele

vation (

m)

Groundwater Withdrawal

NW SE

Slip rate : 1.91 rad/yrLocking Depth: 5.4 km

Figure 4.10: Chaman fault centered profile of line-of-sight velocities from the InSAR rate-map shown inFigure 4.9. Increasing line-of-sight velocities represent motion away from the radar. The gray data areSRTM level 2 3s topography sampled in the same manner as the InSAR data. Larger variances in the to-pographic data indicate larger changes in topography parallel to the Chaman fault. The dashed line cor-responds to the Monte Carlo derived model. Slip rate and locking depth are calculated in the radar line-of-sight. The convex-up feature 25 km northwest of the Chaman Fault corresponds with groundwaterwithdrawal near the town of Qalat, Afghanistan.

71

was acquired and 6 Jan. 2009, that agricultural land-use in the region surrounding the deformation signal

changed.

66˚30' 66˚45' 67˚00'

32˚00'

32˚15'

A

−5 0 5 10

Mean LOS Velocity (mm/yr)20 km

B

66˚45' 67˚00' 67˚15'

Figure 4.11: Comparison of InSAR short-baseline results and Landsat 7 imagery from the Tarnak Rud valleynear the town of Qalat, Afghanistan. A.) Line-of-sight (LOS) rate map of ground subsidence near the townofQalat, Afghanistan (triangle). Positive values indicatemotion away from the radar. Solid arrow indicatesthe flight direction of the satellite and outlined arrow denotes the line-of-sight direction of the satellite.Black triangle marks the location of the town of Qalat and is the same as in Figure 4.9. B.) Landsat 7 imagefrom 18 May 2003 using band combination 4,3,2 to highlight vegetation (red areas).

4.5 Discussion

I havepresented estimates of the interseismic deformation rates at three locations across theChaman

Fault system using space geodetic techniques. Data across the southernmost segment of the Chaman Fault

system indicate that the location of the plate boundary is west of the obvious N-S linear feature mapped

as the plate boundary. My analysis suggests the location of the active plate boundary lies at least 27 km

west of the mud ridge bounding the Las Bela valley and underlies the large Hinglaj Synform (Figures 4.1

and 4.6). Thewide spacing of GPS stations along the Ornach-Nal transect leads to large uncertainty in both

the fault location and the fault locking depth. Assuming that the plate boundary can be described using a

single fault, Monte-Carlo analysis of the GPS velocity transect suggests that the locking depth is nearly 7.4

km with a wide 95% HPD of 1–17.7 km. The location of the fault is similarly uncertain. One value that is

better constrained is the interseismic deformation rate of 14.7mm/yrwith a 95%HPD of 12.8–18.2mm/yr.

72

Analysis of Landsat 7 imagery (Figure 4.6) shows sinistral shearing in the eastern portions of the Hinglaj

synform. Constraining the fault location to the eastern Hinglaj synform, however, fails to significantly

lower the uncertainty in the locking depth and yields similar interseismic deformation rates. Additional

GPS measurements at locations between stations located in Jhal Jhao, Pakistan (ZHAO), and Lak Chuki,

Pakistan (SHFD), will be required to improve estimates of the locking depth and fault location.

Historically, the Ornach-Nal region has been seismically quiet (Zaigham, 1991; Lawrence et al., 1992);

evenMinchin (1907), in his “Gazateer of Las Bela”, notes that this region is not prone to earthquakes. Since

the 1950’s, the ISC catalog has bolstered this view, listing no seismicity near the Ornach-Nal Fault before

1972. During the 1970’s the increasing density of the global seismic network led to an increase in sensitivity

to small magnitude earthquakes, yet the ISC lists no earthquakes larger thanMb 4.9 near the Ornach-Nal

Fault during the period 1972–2010. The general absence of earthquakes combined with their low magni-

tudes and poor station coverage results in very poor depth estimates for these earthquakes. Although the

GPS station density is too low to provide an accurate constraint on the locking depth of the plate boundary

in the Ornach-Nal region, the low seismic productivity combined with the high interseismic deformation

rate shown by the GPS transect (14.7 mm/yr) suggests that the locking depth along the plate boundary is

shallow.

Modeling of Bouger gravity transects between Karachi and the Ornach-Nal Fault by Zaigham (1991)

suggest the presence of east-dipping subducted oceanic crust beneath the southern Kirthar Range. Al-

though depth estimates for earthquakes in the Ornach-Nal region are poor, the presences of sparse, deep

seismicity combined with the interpretation of an east-dipping crustal layer prompted Zaigham (1991)

to suggest on-going subduction of this crustal sliver, termed the Makran-Bela microplate. However, the

low eastward convergence velocity of stations LAKC and SHFD west of Las Bela and the southern Kirthar

Range (∼ 2mm/yr) suggests that subduction of this sliver has ceased. Although the low seismic produc-

tivity within southern Kirthar Range east of Las Bela further suggests that subduction of this sliver is not

ongoing, future GPS occupations between Las Bela and Hyderabad, Pakistan should confirm its cessation.

Farther north, near 31◦N at the town of Chaman, Pakistan (30.89N, 66.51E), I observe an interseis-

mic deformation rate of 7.5 mm/yr across the Chaman Fault. While this deformation rate is low compared

73

with geological estimates of 19–35 mm/yr, the velocity of station CHMC in Chaman, Pakistan relative to

the stable Indian Plate suggests that a total of 19.5 mm/yr of sinistral motion is accommodated by faults

between Chaman and Sukkur, Pakistan. The most likely candidate fault, the Ghazaband Fault, is also a

likely source for the 1935 Quetta earthquake based on relocated epicenters published in Engdahl and Vil-

lasenor (2002). Although the GPS transect crosses the Ghazaband Fault near the town of Pishin, it fails to

reveal any shear across it (Figure 4.8). The transect’s proximity to the Ghazaband Fault’s northern ter-

minus (< 10 km) could account for this observation. Transects further south across the Ghazaband Fault

could help to determine whether it plays a significant role in accommodating plate boundary motion.

The occurrence of the 1892 M 6.5 strike-slip earthquake at the town of Chaman also appears to

conflict with my estimate of a shallow locking depth and low interseismic slip rate along the Chaman

Fault. The description provided by Griesbach (1893) indicates that surface rupture from this earthquake

ran south of the point where the railroad crosses the fault (30.85N 66.52E) for some distance, but did not

extend much farther to the north of the railroad crossing. The transect across the Chaman Fault between

CHMC and SHBG lies 2 km north of this railroad crossing. It is possible that this segment of the Chaman

Fault behaves in a similar fashion to the Parkfield segment of the San Andreas Fault (Lienkaemper et al.,

2006), with the segment north of Chaman locked at depth and accumulating strain at a much higher rate

than the segment near the town of Chaman. If this is analogy holds true, transects across the Chaman

Fault north of the town of Chaman should show progressively deeper locking depths. Indeed, 75 km north

of the town of Chaman, my InSAR rate-map analysis suggests that the locking depth could be deeper than

at the town of Chaman (0–6.6 km near Chaman versus 3–7.8 km near Qalat, Figures 4.8 and 4.10).

Using cumulative seismic moment estimates, previous authors have noted a slip deficit along the

Chaman fault system at latitudes north of approximately 31◦N (Bernard et al., 2000; Ambraseys and Bilham,

2003b). While surface creep along the Chaman Fault has been suggested as an explanation for thismoment

deficit, the best-fit rate-map derived for the region of the Chaman Fault north of the town of Chaman

shows that the fault is not creeping at the surface (Figure 4.10). Given the observed interseismic defor-

mation rate of 16.8 ± 2.7 mm/yr, one would expect Mw 7.0 earthquakes every 60–90 years; this is not

the case as it is likely that the seismic catalog of this region is complete above M 6.5 since at least 1890

74

(Ambraseys and Bilham, 2003b), the absence of earthquakes withmagnitudesM > 7 in the historical record

suggests that the region of the Chaman Fault near Qalat may be due for a large earthquake. Further, this

observation suggests that the low Peak Ground Acceleration prediction for this region for the next 50

years presented in Pakistan Meteorological Department and NORSAR (2007) is too conservative.

In Figure 4.12, I compare the measurements and the maximum sinistral and fault-normal deforma-

tion estimates made from pole-of-rotation parameters published in Altamimi et al. (2007). Estimates along

each fault shown in Figure 4.12 assume that slip is perfectly partitioned across each fault. Since all three

estimates of sinistral motion and both estimates of convergence are lower than the expected velocities

calculated under the assumption of perfect slip partitioning, suggested in each measurement location,

both sinistral slip and convergence are accommodated across multiple structures. In the south, across the

Ornach-Nal Fault, the faults across the Hinglaj Synform appear to accommodate the majority of the sinis-

tral plate motion. Similarly, in the north, across the Chaman Fault near Qalat, Afghanistan, the Chaman

Fault appears to accommodate the majority of the sinistral plate motion. At the latitude of Chaman, Pak-

istan, however, the large deficit in sinistral deformation suggests that the majority of the sinistral plate

motion is accommodated across other faults in the Chaman Fault System.

4.6 Conclusions

I present interseismic deformation rates and locking depth estimates for three locations across

the Chaman Fault System using space geodetic techniques. Along the southern Chaman Fault System,

the Ornach-Nal fault represents the obvious plate bounding fault. Near the town of Las Bela, geodetic

measurements across the fault indicate that the plate boundary is actually located nearly 27 km west

of the Ornach-Nal Fault beneath the Hinglaj Synform. Assuming deformation is accommodated across

a single fault strand, strain is accumulating at a rate of 14.7 mm/yr and is locked to a depth of 7.5 km.

The absence of historical and modern seismicity suggests that significant off-fault deformation west of

the plate boundary must help dissipate accumulated strain. Farther north, at the latitude of Chaman,

Pakistan, the Chaman Fault appears to accommodate as little as 40% of the overall sinistral motion across

the plate boundary with the remaining deformation distributed across faults east of the plate boundary.

75

63˚ 65˚ 67˚ 69˚

26˚

28˚

30˚

32˚

34˚

36˚

Theory: 8.6 mm/yr

Observed: 2 mm/yr

Theory: 8.6 mm/yr

Observed: 5 mm/yr

20 mm/yr

A

Normal Velocity

5 10 15 20 25 30

mm/yr

Sinistral Velocity

B

Figure 4.12: Maximum fault-normal and fault-parallel velocities based on ITRF05 pole-of-rotation loca-tions and rates published in Altamimi et al. (2007) projected along mapped faults on the western boundaryof the Indian Plate. Estimates are derived using the azimuth of the surface trace of plate bounding faultsand assume that slip partitioning is perfect and occurs only along a single fault. A.) Maximum conver-gence estimated assuming perfect partitioning of slip. Locations of convergence observations indicatedby text. B.) Maximum sinistral motion estimated assuming perfect partitioning of slip. Locations of sinis-tral motion estimates indicated by horizontal bars and represent 95% confidence intervals. Note all threemeasurements of sinistral motion and both measurements of fault-normal motion suggest lower ratescompared with perfect slip partitioning.

Geodetic data suggest that the fault is locked at shallow depth (2.7 km) and accumulating strain at a rate of

7.5mm/yr. Although I observeno strain accumulation across thenorthernmost segment of theGhazaband

76

Fault, it is likely that this fault accommodates a large portion of the sinistral motion of the plate boundary

at latitudes south of Quetta, Pakistan. North of the town of Chaman, Pakistan, interseismic deformation

across the Chaman fault is consistent with 16.8 mm/yr of slip beneath a 5.4 km thick locked elastic lid.

Since there have been no recorded earthquakes Mw > 6.5 on this segment of the fault in the past 115

years, it is likely that this segment of the Chaman fault ruptures in earthquakes with magnitudes larger

thanMw 7 with a return interval of> 120 years.

Chapter 5

Fold and thrust partitioning in a contracting fold belt: Insights from the 1931 MachEarthquake in Baluchistan

5.1 Introduction

Between 1931 and 1935 threemajor earthquakes occurred between the Bolan Pass and Quetta in the

Baluchistan province of Pakistan. The first of these, an Mw 6.8 near Sharigh (21:35 UT 24 August 1931),

was followed 66 hours later by the Mach Mw 7.3 earthquake (15:27 UT 27 August). The third and largest

earthquake was the Mw 7.7 30 May 1935 earthquake, that destroyed 90% of Quetta and caused 35,000

deaths (Ambraseys and Bilham, 2003a). In that no similar magnitude earthquakes occurred in the three

decades before or after this sequence, it is very probable that static triggering of these nearby earthquakes

is responsible for their clustering in time.

All three earthquakes liewithin the 150-km-wide zone of deformation between theAsian and Indian

plates, a region bounded to thewest by the Chaman fault and to the east by the Indus plain (Bender andRaza,

1995) (Figure 5.1A). The strike-slip component of slip on the plate boundary is estimated to be 33 mm/yr

from global GPS closure estimates (Apel et al., 2006), and 31 mm/yr from paleomagnetic reconstructions of

the Indian Ocean sea floor (Molnar and Stock, 2009). Geological estimates of slip on the Chaman fault system

(Lawrence et al., 1992) indicate a slip rate of 19–24mm/year in the past 20Myr, and 25–35mm/yr for the past

2 Myr (Beun et al., 1979; Lawrence et al., 1992) suggesting that as much as one third of this shear signal may

be distributed in the fold belt. The plate boundary is regionally transpressive and from seismic-moment

release calculations in the past 200 years, and from the inferred obliquity of the plate boundary to the local

slip vector between India and Asia it has been estimated that strain partitioning results in convergence of

78

the fold belts of up to 13 ± 3 mm/yr (Ambraseys and Bilham, 2003a). This estimate is probably inflated by

the seismic productivity of the past century, which may be abnormally high if the earthquakes presently

under discussion are atypical of long term seismicity. A lower convergence rate is obtained from analog

and numerical modeling of the strain in the region (3–6 mm/yr NW/SE shortening - Haq and Davis, 1997;

Bernard et al., 2000), and this lower rate is consistent with preliminary GPS measurements north of Quetta

presented in this article.

Although triangulation data exist in the region, no remeasurements have been published (Am-

braseys and Bilham, 2003a). However, a first-order spirit leveling line, first measured in 1909 between

Sukkur and Chaman, was remeasured shortly after the 1935 Quetta earthquake (Wilson, 1938). Parts of the

line were raised 65 cm where they crossed the frontal thrusts of the northern Kirthar Range to the west

of Sibi, exceeding the combined errors in the survey by more than an order of magnitude. A preliminary

analysis of these data (Figure 5.2(a)) in the absence of geological constraints, concluded that the asymme-

try in the vertical deformation, if caused by planar slip, was caused by 1–1.2 m of slip on an east-dipping

blind thrust fault between 1 km depth and approximately 25 km depth (Ambraseys and Bilham, 2003a). The

fit between observed surface deformation and synthetic planar slip was appealing, but the asymmetry of

the uplift signal required a counterintuitive easterly dip to the frontal thrust. The description of a wedge-

thrust geometry with an east-dipping shallow ramp at this location in the literature, however, appeared

to confirm its presence (Banks and Warburton, 1986). Yet, when the surface deformation associated with

slip on the two faults of their triangle zone were examined in detail by Garcia et al. (2006), no combina-

tion of slip was found that resulted in an improved fit to the leveling data (Figure 5.2(b)). Models with

more complex geometries also failed to improve the fit. The two previous interpretations, based as they

were on limited structural information, may now be discarded due to the availability of seismic reflection

data from the Sibi and Mach areas, controlled by stratigraphic information from numerous boreholes.

I present structural interpretations in the next section that demonstrate that the frontal thrusts of the

Kirthar range west of Sibi indeed dip to the west (Bannert et al., 1992; Schelling, 1999a).

79

66˚ 68˚28˚

30˚

Sharigh1931

Quetta1935

Mach 1931

A

100 km

66˚ 68˚

leve

ling lin

e

chmc

qila

sarn surb

kach

qtag

shbg

20 mm/yr

Sibi

Quetta

Kalat

Mach

Chaman

B

100 km

Figure 5.1: A.) Recent seismicity (Mw > 5) and instrumental locations for the Sharigh, Mach (stars) andQuetta earthquakes (focal mechanism beachball) and their inferred causal faults (Quetta rupture dashedand Bannh fault shown as surface thrust NE of the instrumental epicenter). Focal mechanisms scaled ac-cording to magnitude - the largest focal mechanism is Mw 7.7 (Singh and Gupta, 1980) and the smallest isMw 5 (all from the Global CMT). B.) Interpolated Intensity VIII isoseismals for the three earthquakes, thepath of the 1909–1936 leveling line and GPS velocity vectors 2005–8 relative to fixed India. The approxi-mate rupture zone of the Mach earthquake is shown by the rectangle. The intensity-derived epicentersare shown on eachmap as a star. The Quetta centroid solution lies at the opposite end of the rupture fromthe intensity solution as a result of directivity.

Dis

plac

emen

t (m

m) a

Dep

th (k

m)

b

Distance (km)

c

Figure 5.2: Schematic sections of vertical deformation and subsurface geometry of previous attempts toemulate observed uplift data in the Mach earthquake (Figures 5.2(a) and 5.2(b)). These models invokeduniform subsurface slip on shallow east-dipping planar thrusts. In Figure 5.2(a) planar, uniform slip isinvoked with no structural control (Ambraseys and Bilham, 2003a). In Figure 5.2(b) the speculative wedgethrust geometry of Banks and Warburton (1986) constrains two fault planes on which combinations of uni-form slip were imposed to obtain the best-fitting surface uplift (Garcia et al., 2006). Spatially variable slipon the west-dipping Bannh fault (Bannert et al., 1992; Schelling, 1999a) is presented here (2c).

80

5.2 Structural setting of the Bolan Pass Region

TheBolanPass region of Baluchistan (Figure 5.3) is located along the deformation front of thenorth-

ern Kirthar Range, where north-south trending structural systems (folds and thrust faults) of the Kirthar

Range give way to the more complexly oriented structural systems of the Quetta Syntaxis. Detailed struc-

tural field work carried out in the Bolan Pass region and elsewhere in the Kirthar Range during the late

1990’s (Schelling, 1999a) indicates that the deformation front of the northern Kirthar Range is dominated

by east-vergent, contractional fold-fault systems that give way to strike-slip oriented fault systems along

and to the west of the Quetta Plateau. In addition, surface structural data has allowed the geometries and

orientations of fold-fault pairs to be defined across and along the mountain front at different tectono-

stratigraphic levels in the Bolan Pass region. Surface structural geometries from the Bolan Pass have been

projected to depth, and in conjunction with interpreted seismic data from the Bolan Pass and adjacent

Sibi Trough areas, a balanced structural cross section has been constructed across the Bolan Pass in the

vicinity of the leveling line examined in this paper (Figure 5.4).

As shown on the cross section of Figure 5.4, the frontal structural system of the Bolan Pass area

is defined by an east-vergent, asymmetric fault-propagation fold (the Dezghat Anticline) and underlying,

west-dipping thrust fault system that is known from seismic data to flatten near the base of the Siwalik

Group, a roughly 6 km thick stratigraphic section of Miocene-Pliocene sandstone, shale, and conglomer-

ate that define foreland basin fill to the actively subsiding Sibi Trough (Indus Basin). Tectonic shortening

across the Dezghat Anticline and associated thrust faults is on the order of several kilometers (Figure 5.4).

However, all of the structural uplifts identified to the west of the Dezghat Anticline are known from sur-

face and subsurface structural data to involve the Jurassic Chiltan Limestone and overlying Cretaceous

through Eocene stratigraphic section, including the Goru and Sembar formations, the Parh and Dung-

han limestones, the Ghazij Shale, and the Kirthar Limestone (Figure 5.4). Structural uplift of the Chiltan

Limestone, between the Parhi Jhal Anticline mapped to the west of the Bolan Pass, and the Sibi Trough

located to the east of the Kirthar Range, is on the order of 9 to 10 km (Figure 5.4), and associated tectonic

shortening across the same structural systems, as determined from surface and subsurface data, is esti-

81

Bannh #1

Ghazij Formation

Dunghan FormationSiwalik Group

Quaternary Alluvium

Kirthar Formation Cretaceous

Jurassic Chiltan Limestone

Leveling Line

Thrust Fault

Q

TK

Td

K

TKTK

TKTKTKTK

TKTK

TKTK

TKTK

TKTK

TKTK

TdTd

TdTd

TdTd

TdTd

TdTd

TdTdTdTd

KK

KK

KK

KK

KK

KK

QQ

QQ

QQ

QQ

QQ

QQ

QQ

QQ

QQ

QQ

QQ

QQ

QQ

QQ

QQ

QQ

Cross Section (Figure 4)

Negh

rain

i Ant

iclin

eDar

kin

Ant

iclin

e

Parri

Jhal

Ant

iclin

e

Dezga

t Anti

cline

Bannh

Anti

cline

QUETTAPLATEAU

SardarKhel

Bolan Pass Anticline

Negh

rain

i Ant

iclin

eDar

kin

Ant

iclin

e

Parri

Jhal

Ant

iclin

e

Cross Section (Figure 4)

Mach

Dezga

t Anti

cline

Bannh

Anti

cline

SIBITROUGH

DadharD

adha

r Syn

clin

e

Figure 5.3: Geological map of the Bolan Pass region of the northern Kirthar Range, showing the locationsof the balanced structural cross section and leveling line discussed in the paper.

82

MSL

3 km

-5 km

-10 km

-15 km

3 km

MSL

-5 km

-10 km

-15 km

DEZGHATANTICLINE

BOLAN RIVERANTICLINE

PARRI JHALANTICLINE

QUETTAPLATEAU

KUMBRI NALASYNCLINE

Bolan River Valley

Pre-MZ

Pre-MZ

Pre-MZ Tg Td

K

Tg

J

Tdc

Tu TuTm Tm

TslTsl

T

J

J

T

TTdc

Cretaceous

Dunghan Fm.

K

Td

Upper Siwaliks

Dadhar Cglo.

Tu

Tdc Pre-MesozoicPre-MZKirthar Fm.

Ghazij Fm.

Tk

Tg

Chiltan Ls.

Triassic

J

TLower SiwaliksTsl

Tm Middle Siwalks

Figure 5.4: Balanced structural cross section across the deformation front of the northern Kirthar Rangein the Bolan Pass area and in the vicinity of the leveling line. See Figure 3 for cross section location andtext for discussion.

mated at approximately 15 km. This requires that thrust faults exposed at the surface and that involve

the Chiltan Limestone are relatively high-angle structural features, with measured, near surface dips of

roughly 60 degrees and estimated fault angles of 30 to 45 degrees at 10 to 20 km depth (Figure 5.4). In

addition, 9 to 10 kilometers of uplift across the combined Bolan Pass and Parhi Jhal Anticlines requires a

mid-crustal decollement surface at a depth of 18 to 20 kilometers beneath the Quetta Plateau, as indicated

on the cross section of Figure 5.4. This 18 to 20 km decollement depth is well below the projected, base-

Triassic stratigraphic level of known lithology from the Kirthar and nearby Sulaiman mountain ranges,

and therefore the lithology at the basal decollement level beneath the Quetta Plateau remains unknown,

and may actually be located in basement rocks.

Additional decollement surfaces of the Kirthar Range have been identifiedwithin the Eocene Ghazij

Shale, which separates the underlying, competent carbonates of the Chiltan-Dunghan limestones from

the overlying Kirthar Limestone and Siwalik Group (Figure 5.4). Significant deformation associated with

these Ghazij Shale decollement surfaces is restricted to an area above the sub-surface, frontal fault-ramp

identified beneath the Bolan Pass, where the underlying (basal) decollement surface to the Kirthar Range

climbs from a depth of 18 kilometers or more beneath the Parhi Jhal and Bolan Pass Anticlines to the up-

per, basal Siwalik Group decollement surface identified beneath the Dezghat Anticline at roughly 6 km

depth. Thrust faults originating within the Ghazij Shale result in short- (several hundred meter) wave-

length anticline-syncline pairs and the development of exposed back-thrust surfaces along the east-limb

83

of the Bolan River Anticline. These latter fold-fault systems have accommodated less than 1 km of tec-

tonic shortening, though as exposed structural systems in the Bolan Pass area there is little question that

thrust faults originating within the Ghazij Shale will affect surface deformation across the Bolan Pass area,

as indicated from the leveling data discussed in this paper.

5.3 GPS measurements of convergence and shear between the Asian and Indian Plates

GPS measurements in Pakistan are historically of limited coverage and duration. Campaign mea-

surements from six sites with locations between the town of Chaman, 30 km west of the Chaman fault,

and the town of Kach, approximately 70 km NE of Quetta, have been measured at least twice in the period

2006–2008, and compared to continuous measurements made in Karachi (not shown) and Quetta (qtag)

(Figure 5.1B). The continuous points in Pakistan are operated from flat-roofed concrete frame buildings

and the campaign points are measured on bipods set on stainless steel screws cemented into exposed

rock. GPS observations were recorded either with Trimble NetRS, 5700 or R7 receivers using a 30 second

sampling rate, and processed using an elevation cut-off angle of 10 degrees. Campaign data have dura-

tions of 3–7 days from each site. The daily data from these sites were processed along with data from 10

regional IGS stations using GAMIT version 10.34 (King and Bock, 2002). The regional solutions were then

combined with global solutions from SOPAC (ftp://garner.ucsd.edu/pub/hfiles) using GLOBK/GLORG

version 5.16 (Herring, 2002) to determine time series and velocities in the ITRF2005 reference frame. These

velocities were then transformed into an Indian Plate-fixed reference frame using pole of rotation param-

eters determined by Bettinelli et al. (2006).

The processed campaign GPS data are associated with formal uncertainties of ±3 mm/yr, and the

continuous data with uncertainties of ±1 mm/yr (Figure 5.5). Not shown on the figure are the velocity

vectors for Karachi and Nagar Parkar (north of the Bhuj 2001 earthquake) that move at approximately

the velocity of the Indian plate, suggesting that little deformation occurs across the Indus delta, or near

the Bhuj region. The SSW velocity of Quetta is anomalous relative to the points on the east-west transect

through Chaman due to intense groundwater withdrawl (S. Khan, Univ. Houston, personal communica-

tion 2010). The most easterly point on the traverse, Kach, is also anomalous in that it shows no conver-

84

−5

0

5

10

West V

elo

city (

mm

/yr)

−50 0 50 100

Distance (km)

CHMC SHBG

QILA

SARN LAKP

SURB

QTAG

KACH

Chaman Fault

West East

Figure 5.5: GPS velocities projected E-W showingwestward velocities relative to stable India. For locations,see Figure 5.1B. The GPS points, with one exception, show convergence with fixed India at 5± 1mm/yr.The one exception is QTAG, the continuous GPS point at Quetta.

gence with India. I discuss this observation in Chapter 6

Thus the limited view of motions within this complex region of shear and convergence afforded by

the GPS data permit only the simplest of interpretations at present. In Figure 5.5 I present the Chaman

Fault-perpendicular velocity profile; in Chapter 4, I analyze the Chaman Fault-parallel velocity profile.

I interpret the mean westward translation of five of the seven GPS points depicted in Figures 5.1

and 5.5 towards the Indian plate at 5 ± 1 mm/yr as indicative of the transpressional convergence of the

Sulaiman fold belt, which I use as a proxy for maximum convergence rates in the Kirthar range. The GPS

measurements were obtained north of Quetta at approximately 30◦N, where the fold belt is significantly

wider than at the latitude of the Mach earthquake. The lower transpressional obliquity of the Chaman

system south of 30◦N, and its narrower width, suggests that convergence occurs there at a lower rate. I

cannot as yet quantify this from direct observation, but I assume that the rate is at least half that of the

rate measured north of Quetta, i.e. current east-west convergence near the epicenters of the Quetta and

Mach earthquakes is probably 2.5 to 5 mm/year.

85

5.4 Macroseismic location of the Mach earthquake

The Mach earthquake is named for the railway headquarters at Mach that were heavily damaged

in the earthquake (Ambraseys and Bilham, 2003a). The jail was destroyed and 400 prisoners were briefly at

large. Although no surface rupture was recorded, the parapets of a 140 m long approximately E-W bridge

converged 20 cmwithout being tilted. Numerous rock falls occurred at the time of the earthquake, raising

clouds of dust near Mach and the Bolan Pass to the SE.

Although instrumental locations and magnitudes are available for the 1931 and 1935 earthquakes

(Figure 5.1 and Table 5.1), additional knowledge of the extent and azimuth of their rupture zones can

be inferred from intensity data recorded for each event. The data are available in the form of damage

reports to structures near their epicenters, and from felt reports at larger distances. Previous analysis of

the intensity data for the Sharigh, Mach and Quetta earthquakes have interpolated isoseismal contours

for a range of intensities to determine the most probable location of their rupture zones. Banana-shaped

isoseismals drawn by West (1934) for the highest intensities are distorted by the uneven coverage of his

data. Ambraseys and Bilham (2003a) re-evaluated these data supplemented by additional observations and

concluded that the highest isoseismals for the second two earthquakes form north-elongated polygons

(Figure 5.1A). Insufficient data for the Sharigh earthquake were available to form definitive conclusions,

except that its epicentral location was close to the town of that name.

Table 5.1: Instrumental and inferred macroseismic locations for the three earthquakes.

Event Date Instrumental Mw Minimum Magnitude Mi Minimum VarianceEpicenter Epicenter Epicenter

Sharigh 24 Aug. 1931 31.1N 67.7E 6.8 29.87N 67.62E 5.9 30.12N 67.60EMach 27 Aug. 1931 29.9N 67.6E 7.3 29.55N 67.55E 7.2 29.22N 67.47EQuetta 30 May 1935 28.87N 66.4E 7.7 30.10N 66.92E 7.6 30.18N 66.92E

I have subjected these same observed intensities to a more rigorous analysis using the methods of

Bakun and Wentworth (1997). This approach does not contour isoseismals but instead contours the most

probable locations for the epicenter using a grid search and assumptions about attenuation of shaking in-

tensity with distance. The method uses recent earthquakes from elsewhere in the region for which loca-

86

tion andmagnitude are known and for which intensity data are also available, to quantify the attenuation

of shaking intensity with distance (Chapter 3). The method then calculates the most likely magnitude for

the earthquake, were it to have occurred at points on a hypothetical regularly spaced grid centered on the

centroid of maximum intensities. The resulting values on the grid are then contoured to provide a series

of iso-magnitude contours surrounding a closed contour of minimummagnitude for the earthquake. The

center of this minimum contour is the epicentral location of the smallest possible earthquake that could

have caused the observed intensity distribution (Figure 5.6 and Table 5.1). If observations are noise-free,

well distributed in intensity range, and endowed with good azimuthal coverage, these contours tend to be

elliptical or circular, however, if the coverage is azimuthally poor, or of low quality the resulting contours

may be complex with multiple minima.

67˚ 69˚28˚

30˚

67˚ 69˚ 67˚ 69˚

6.4

6.66.8

6.8

Sharigh

7.2

7.4

Mach

7.8

8

Quetta

Figure 5.6: Macroseismc epicenters for the Sharigh, Mach and Quetta earthquakes. The dashed contoursin this figure are not isoseismals but iso-magnitude contours using the method of Bakun and Wentworth(1997). They indicate the required magnitude for each earthquake had it been located on these contours.The preferred macroseismic epicentral location lies within the closed contour of the minimum-variancesolution shown as solid lines while the stars represent the instrumentally located epicenters.

The magnitude contours provide no estimate of the variance between magnitudes predicted from

each observation at each point in the grid search. A new set of contours is generated based on the vari-

ance of the magnitude estimates derived for each point on the grid were the earthquake to have occurred

at that point. When these variances are contoured, a region of minimum variance is obtained, usually,

87

but not always, close to the minimum magnitude solution. For the three earthquakes in Baluchistan the

minimummagnitude locations are indeed close to their minimum variance locations, typical of an accept-

able macroseismic solution (Table 5.1). The minimum variance contours are assigned probabilities, with

the most probable location for the epicenter being the location of minimum variance. The intersection of

the minimum variance location with a contour of the iso-magnitude solution indicates the most probable

magnitude for the earthquake.

The solutions for all three earthquakes are listed inTable 5.1 and shown in Figure 5.6. The absence of

complexity to theminimummagnitude contours and their coincidencewith themarginallymore complex

minimum variance locations in each case is notable. The most probable location for the Mach epicenter

lies near the southern end of its rupture zone and near its instrumental location. I ignore the smallermini-

mum noted north of Sibi. In contrast, the location of the minimum variance epicenter for the 1935 Quetta

earthquake lies near Quetta, more than 130 km NNE of the instrumental location for the earthquake. I

assume that the high intensities reported from Quetta were enhanced by directivity in the direction of

rupture propagation (Singh and Gupta, 1980; Day et al., 2008). The minimum-variance epicentral location

for the Sharigh event is found to be approximately to the north of the Mach event, west of the epicentral

location inferred by earlier investigations, and in Chapter 6, I discuss candidate causal faults.

Using the constants derived in Bakun andWentworth (1997) for California I infer theMachmagnitude

to have beenMw 7.2 in good agreementwith the value ofMw 7.3 derived byAmbraseys and Bilham (2003a).

The inferred magnitude for the Sharigh event is significantly smaller, Mw 5.9 instead of 6.8, probably

the result of the sparse sampling of macroseismic data points for this event. The preferred location for

the Sharigh event lies to the west of previously inferred locations, but no causal fault can be identified

there from geological evidence or microseismicity. Recent CMT solutions (Dziewonski et al., 1999) in the

area (Figure 5.1A) indicate that the newly located Sharigh event lies in a transition region between NW-

SE directed compression in the Bolan Pass region and N-S compression to the north. Focal mechanisms

nearest to the epicenter of the Sharigh earthquake show a combination of NE dipping thrust faulting and

thrusting along N-S directed decollements.

88

5.5 Leveling data

The Sukkur-Quetta leveling line (Figures 5.1B and 5.7) was originally surveyed between 1909 and

1914with a benchmark spacing of 2–5 km (Wilson, 1938; Ambraseys and Bilham, 2003a). The leveling data are

associated with random errors that growwith the square root of the distance traversed (L km) as k√Lmm

where k = 0.65, a constant derived from circuit closure errors in India (Lenox-Conyngham, 1916). In addition,

a systematic height-dependent error is present in the data that is typically less than 1×10−6 of the height

above the starting point of the line in kilometers. The 65 cm of vertical deformation near the Bolan Pass

exceeds both systematic and random errors in the data by more than an order of magnitude. The leveling

bench marks were geo-referenced from 1”= 1 mile topographic sheets and have resulting uncertainties

of up to 30 m (listed in Ambraseys and Bilham, 2003a). I project these irregularly spaced leveling data at

N110◦E along a line perpendicular to the trend of folding in the Kirthar range (Figure 5.7).

No surface rupture was reported for the 1931 Mach earthquake, although it is possible that West’s

post seismic investigations did not include traverses across the frontal thrusts of the Kirthar range except

near Sibi (West, 1934). To proceed with the analysis of the leveling data, I assume that slip on one or

more of the mapped subsurface faults was responsible for the observed uplift. I digitized these subsurface

faults from the geological cross-sections, forming curved fault segments from a series of contiguous, 3-

km-wide, planar segments. I then permitted various combinations of contiguous segments to slip. Each

segment with non-zero slip contributes to the surface deformation field (Okada, 1992) and I searched for

smooth distributions of slip on contiguous segments that most precisely produced the observed surface

deformation. The analytical procedure I adopted was to shift the surface projection of the parameterized

faults relative to the leveling data, and invert for slip using the Green’s functions for each segment. By

minimizing the sum of squared residuals, I determined the optimal offset between the leveling data and

the modeled geometry.

The misfit between the parameterized faults and the projected leveling data is minimized with

slip on deep segments of the Dezgat Thrust with contiguous slip on segments of the Bannh Fault which

branches from it eastwards towards the surface (Figure 5.7). Maximum observed slip of 3.2 m occurs in

89

-50 0 50

verti

cal d

ispl

acem

ent

mm

, 193

6-19

09

Cha

man

f.

Que

tta

+ - + -subs

urfa

ce s

truct

ure

& m

ean

topo

grap

hy, k

m

1

2

0-10

100 km

Sibi

600

400

200

0

1931 Mach Mw 7.3 rupture

Mw

7.5

synthetic fit

observed

SE distance from Quetta

Dezghat F.Bannh F.

Figure 5.7: Leveling data, topographic relief and subsurface section simplified from Figure 5.4. The syn-thetic fit to the data results from spatially varying slip on the Dezghat and Bannh faults (dashed line onsection).

90

segments between 7 and 5 km depth, up-dip from an inferred interseismic locking line at 8 or 9 km depth,

where I observe minimum slip. The location of inferred interseismic locking is not constrained by ob-

servations of interseismic deformation, and hence there exists some uncertainty as to its true location.

However, in Section 5.6 I undertake additional numerical models that are consistent with this identifi-

cation. A deeper locking line would decrease the inferred along strike length of the Mach rupture. The

total width of the rupture above the 9-km-deep locking line is 42 km and themean slip is 1.2 m (Table 5.2).

Assuming a seismic moment of 1.1× 1027 dyne-cm, and that all the slip occurred seismically, the best-fit

slip distribution requires an along-strike rupture length of 72.2 km.

A broader region of uplift near Quetta (Figure 5.7) can be explained by invoking minor slip on an

uneven decollement or by invoking slip on one of several mapped listric faults near there, either before or

after the 1931 earthquakes. I note that the fit to these minor regions of uplift in the data are non-unique

since they are not constrained by well-defined subsurface geometry.

5.6 Discussion: the earthquake cycle in a ramp-flat-ramp system

The vertical displacement data fit in the foregoing section includes all deformation that occurred

between 1909 and 1936. Thus the data include not only co-seismic slip but possible post-seismic slip, if any

occurred. I show that this is likely with a series of elastic models that assume perfectly frictionless slip

below a locking line, and rupture at shallower depths (Figure 5.8). The models take the form of those de-

scribed by Feldl and Bilham (2006) inwhich a series of contiguous, frictionless boundary elements are driven

along a complex rupture surface in an uniform elastic half-space by a far-field displacement imposed at

depth. The geometry of this selected far-field driving condition is not critical to the models. Similar re-

sults are obtained by imposing regional contraction on the fault system, or imposed remote thrusting.

The boundary element computation calculates the amount of slip required on contiguous elements to

minimize stress in their vicinity.

The results from three calculations (Table 5.2 and Figure 5.8) illustrate models that emulate pre-

seismic, coseismic slip, and finally, slip assuming no interseismic locking below the region of coseismic

rupture (upper line). For the coseismic and preseismic slip calculations, the slip distribution is calculated

91

Table 5.2: Observed (“Obs”) and synthetic slip on the decollement. Segments are free to slip in responseto 10 m of thrust displacement imposed on the deepest fault segment, a value scaled to approximate themean observed coseismic slip. “Co-8” refers to coseismic slip shallower than approximately 8 km depth,and “Co-9” refers to coseismic slip from one segment deeper at approximately 9 km depth. “No-Lock”indicates the slip that would occur in the absence of interseismic locking, and “interseismic” indicates thesynthetic slip that occurs below a locking line at 9 km depth. I have scaled the driving element to 10 kmso that that synthetic slip approximates the mean slip derived from the observed leveling data.

Distance Depth Length Dip Obs No-Lock Co-8 Co-9 Interseismic(km) (km) (km) (m) (m) (m) (m) (m)104 -3.5 3.03 7.9 0.28 0.86 0.47 0.463 0101 -3.92 3.01 5.14 0.35 1.34 0.72 0.713 098 -4.19 3.01 3.39 0.3 1.71 0.9 0.901 095 -4.36 3 2.51 0.36 2.03 1.06 1.06 092 -4.5 3 2.47 0.35 2.32 1.21 1.21 089 -4.62 3 3.28 0.4 2.62 1.34 1.35 086 -4.8 3.01 4.96 0.76 2.91 1.48 1.5 083 -5.06 3.03 7.63 1.17 3.2 1.61 1.63 080 -5.46 3.01 4.33 1.21 3.42 1.69 1.73 077 -5.69 3.01 3.83 1.63 3.6 1.73 1.79 074 -5.89 3.01 3.68 2.49 3.75 1.73 1.83 071 -6.08 3.03 7.77 3.16 3.93 1.7 1.84 068 -6.49 3.03 12.79 2.86 4.2 1.59 1.81 0

65.05 -7.16 3.02 28.75 1.56 4.64 1.31 1.71 062.4 -8.61 3 29.88 0.68 4.96 0 1.37 059.8 -10.11 3.02 12.13 0.54 5.29 0 0 1.656.85 -10.74 3.01 5.68 0.84 5.68 0 0 2.4553.85 -11.04 3.01 4.22 1.04 6.04 0 0 3.1650.85 -11.26 3.01 3.63 0.9 6.35 0 0 3.847.85 -11.45 3 2.38 0.68 6.65 0 0 4.3944.85 -11.58 3 1.96 0.65 6.95 0 0 4.9741.85 -11.68 3 1.71 0.66 7.27 0 0 5.5638.85 -11.77 3 1.49 0.42 7.61 0 0 6.1635.85 -11.85 3 1.29 0.09 7.99 0 0 6.8132.85 -11.91 3 1.11 0.33 8.43 0 0 7.5329.85 -11.97 3 0.96 0.96 8.98 0 0 8.4126.85 -12.02 400.02 0.57 0.48 10 0 0 10

92

by assuming the decollement surface is frictionless and free to slip in response to 10 m of convergence

applied from the west, either for the entire fault surface (no-locking) or up to a locking line at 8 or 9 km

depth (pre-seismic). The coseismic calculation corresponds to seismic slip above the locking line, and the

input to this model is the static strain field developed from the preseismic slip distribution determined

from the interseismic slip calculation (Table 5.2).

The selection of a 10mdrivingdisplacement is arbitrary because the synthetic output scales linearly

with input. However, I note that this input value results in synthetic co-seismic slip that approximates

the coseismic slip inferred from the leveling data. The 10 m input condition corresponds to 2000 years of

convergence at 5mm/year. The selection of the locking line, the transition between downdip interseismic

creep and the locked seismogenic rupture zone, was investigated by running models with incrementally

increased locking depths and by examining the resulting slip distribution with that inferred from the lev-

eling data. The best fitting coseismic slip distribution occurs where interseismic locking occurred above 8

or 9 km depth. The coseismic slip distributions resulting from locking at each of the 8 km and 9 km depths

are listed in Table 5.2.

The spatial distributions of observed and synthetic co-seismic slip show similarities. Peak slip in

synthetic and observed data coincides in up-dip location but the ratio of peak slip to average slip is less in

the numerical experiments than observed in the Mach earthquake. Observed slip is twice the synthetic

slip at 6 km depth, and half the synthetic slip at 4.5 km depth. No simple changes in fault geometry, or

freely slipping width were able to emulate the localized maximum slip at 6 km depth.

The observed minimum in slip that occurs at the inferred locking line is of special interest. A sig-

nificant slip deficit (3 m) occurs here as a result of preseismic and postseismic pinning at the locking line.

The resulting slip deficit is analogous to the slip deficit that occurs between two contiguous strike-slip, or

normal faults that slip sequentially. Manighetti et al. (2005) suggests that the stresses generated by this slip

deficit are released in off-fault deformation through the creation of secondary faults and folding.

A paradox, however, arises in the thrust fault I am considering, for if the locking-line is pinned

over several earthquake cycles, the hanging wall cannot advance over the footwall. Thus although the

minimum slip in the model is confirmed by the 1909–1936 leveling data, slip may occur at the locking line

93

-12

-10

-8

-6

-4

-2

0

depth (km)

(no locking)

synthetic pre-seismic synthetic

co-seismic

observed co-seismic

locking line

rupturedecollement

8 1012

slippinned

driven element

imposed 10m slip

West East

Dezghat Fault Bannh F.

slip

def

icit

displa

ceme

t (m)

2

4

6

8

10

Figure 5.8: Geometry of the active decollement and frontal thrust (bold line with depth scale right), andinferred slip on segments shallower than 9 km (grey envelope) compared to synthetic slip (slip scale left).The calculated slip for the entire fault is given by the top staircase-line (21, 3-km-long freely-slippingsegments responding to an input displacement of 10m imposed from the left (west)). The lower staircase-lines are formed from two calculations: slip anticipated below a locking line at 9 km depth (syntheticpre-seismic slip), and the slip during rupture at shallower depths that occurs when this interseismic slipdistribution drives co-seismic rupture (synthetic co-seismic). The difference between the two lower stair-case lines and the upper staircase is the slip deficit caused by interseismic locking at 9 km depth.

94

at times not sampled by these data. Stress conditions for slip in the region (afterslip) aremost favorable for

this translation shortly after the earthquake before the shallow fault “heals”, yet a significant slip deficit

remains four years after the earthquake. Thus, if slip occurs here, it must do so over a period of many

decades after each earthquake.

Alternatively, if the locking line is truly locked for numerous earthquake cycles, the stored elastic

energy there must eventually be released in a much larger earthquake. The significant variability in slip I

see in the Mach earthquake could in fact be the result of the 1931 event being driven partly by elastic en-

ergy stored from a previous earthquake cycle. If this occurred it would suggest that previous earthquakes

terminated at a shallower locking line than that I infer for 1931. There is some evidence to suggest that

the Himalayamay exhibit similar enigmatic behavior, withmost decollement earthquakes associatedwith

3–7m of slip and no prominent surface rupture, but with infrequent earthquakes causing surface ruptures

with slip of as much as 24 m (Feldl and Bilham, 2006; Bilham and Szeliga, 2008).

I section 5.7 I discuss the implied discrepancy between the amount of convergence (approximately

10 m) required to drive 3 m of coseismic slip of the frontal thrust.

5.7 Geodetic convergence, slip potential and renewal time

In many paleoseismic estimates of earthquake recurrence interval, the renewal time for an earth-

quake is estimated from the strain rate applied to a fault, a number that is derived from the present-day

geodetic displacement rate measured in the region. Thus one might anticipate that a 3 mm/yr conver-

gence rate applied to a fold-and thrust-belt would permit earthquakes with 3 m of slip every thousand

years. The perfectly elastic frictionless calculation indicates that the renewal of the Mach earthquake us-

ing this approach would err by a factor of 8, because only one-eighth of the convergence is manifest as

slip on the frontal thrust (Table 5.3). The remaining convergence is presumably accommodated by folding

and thickening of the fold and thrust belt.

The effect occurs because the ramp separating the deepdecollement from the shallow frontal thrust

acts as a buttress to motion. The ratio of input displacement (geodetic convergence) to frontal fault slip

(measured coseismic slip) depends on the ratio of the depth of the decollement to the mean depth of the

95Table 5.3: Calculations of partitioned convergence. Geometric relations between applied geodetic dis-placement and slip on the Dezghat/Bannh thrust fault for a range of hypothetical decollment depths (theactual depth is believed to lie in the range 18–20 km). The imposed displacement, S, is that calculated tocause themeanobserved coseismic slip, s, in theMach earthquake. D is themeandepth of the decollement,and d, is the approximate starting depth of the frontal thrust above a steeper ramp connecting the two.The ratio S/s is a proxy for the increase in the recurrence interval for earthquakes on the frontal thrustscompared to the time that would be calculated from geodetic convergence rates of the entire range.

Depth D (km) Imposed S (m) ratio D/d ratio S/s6 6 1 28 8.9 1.3 310 12 1.7 412 15.75 2 5.218 23.5 2.9 8

frontal thrust. As discussed in Section 5.2, the depth of the decollement may lie at 18–20 km depth. In Fig-

ure 5.8 the depth of the decollement, D, is placed at 12 km depth and the depth of the flat is approximately

6 km, a ratio of approximately 2. From a suite of numerical models in which I varied the decollement

depth while maintaining the shallow geometry (Table 5.3) I derive the following relationship:

C = .00154D − 3.2m,

where C is the (geodetic) convergence of the entire range and D is the depth of the decollement, both

measured in meters. The constants in the equation would be modified in systems with different shallow

thrust geometries, but my basic finding would be unaltered.

I find that if the decollement lies at 6 km, the renewal time is approximately doubled (Table 5.3),

and if it lies at 10 km the renewal time is quadrupled. Only for the case where the system is a simple

planar ramp does the renewal time obey a simple 1:1 relationship between slip-potential and geodetic

convergence. Note that these calculations are for an infinitely long fault. By reducing the along-strike

length of the rupture the ratio of convergence to potential slip is increased yet further, as was found for

the synthetic scaling law for the Himalaya (Feldl and Bilham, 2006). If I use the 18–20 km depth inferred for

the depth of the decollement underlying the northern Kirthar range, the Dezghat thrust “receives” only

12% of the convergence applied to the entire Kirthar range between the Chaman Fault and the plains of

the Indus River.

96

How can one explain this significant discrepancy? It would appear that the ramp acts as a buttress

to sedimentary layers driven from the west. Since convergence is not released as slip on the shallower

fault above and to the east of the ramp, it must be manifest as thickening of the sediment pile to the west.

The mean topography west of the ramp is 1.5 km higher than to the east of the ramp, and the seismic

section is shortened by approximately 15 km. East of the ramp the frontal Dezgat/Bannh fault system has

been shortened by less than 1 km.

5.8 Sequential triggering of ruptures

The three earthquakes are unusual in that most of the seismic moment release in the Baluchistan

region in the last 150 years occurred within the four years following the first of these earthquakes (Figure

5.9). The clustering of these three large Baluchistan earthquakes near 30◦N has a low probability of occur-

ring by chance; hence some form of triggering appears probable. The mechanisms of stress transfer are

currently speculative, especially for the first two events that occurred within 66 hours of each other. In

Chapter 6, I discuss possible mechanisms for the Sharigh earthquake by utilizing both its location within

the Quetta Syntaxis Shear Zone and its temporal relationship with the Mach earthquake. However, the

geometric relationship between the Mach and Quetta earthquakes renders a causal link substantially eas-

ier to comprehend. Rupture of the Dezgat thrust in the Mach earthquake reduced compressive east-west

stresses and acted, in a sense, to unclamp the fault-normal stress on the Quetta strike-slip fault 60 km to

the west. The instantaneous Coulomb failure change at Quetta is calculated to be 10 to 70 kPa, depending

on the nucleation depth of the Quetta earthquake, a stress change that ismore than sufficient to trigger an

earthquake (Stein et al., 1994). However, these instantaneous stress changes were apparently unimportant

because the earthquake was delayed bymore than three years. If the occurrence of the Quetta earthquake

less than 4 years after the Mach earthquake is not a coincidence, then some form of stress diffusion, or

viscous creep is required.

While poroelastic or viscoelastic processes in the body of the fold and thrust belt, or below it (Freed,

2005) are adequate to cause the observed delay, I consider here an alternative mechanism - viscous creep

on the decollement surface. The rate of propagation of the deformation front between Mach and Quetta,

97

1e+24

1e+25

1e+26

1e+27

1e+28

1e+29

Mo

ment (d

yne−

cm

)

1800 1850 1900 1950 2000

Year

1 Fault Length2 Fault Lengths3 Fault Lengths

Figure 5.9: Space-time history of seismic moment release as a function of distance from the inferredMach1931 earthquake rupture zone. More than 89% of the total seismic moment release in the past 200 years(within a radius of 500 km centered on the Mach earthquake) occurred between 1931 and 1935. All knownearthquakes larger than M6.5 are included in this plot.

98

had it occurred linearly, is approximately 18 km/yr. I note that 4 years after the earthquake I infer afterslip

10 km below the locking line to have amounted to approximately 1 m, with approximately 0.5 m of slip 20

km below the locking line. If this decay rate continued linearly downdip towards the Quetta fault, the slip

on the decollementmay have amounted to 10-20 cmnear Quetta by 1935. The leveling data are insensitive

to slip on a planar, sub-horizontal surface, but the bulge in the data east of Quetta (Figure 5.7) suggests

that slip of some form occurred 40-50 km west from the Mach event. Slip may also have occurred prior

to the Mach event in the region between Quetta and Mach, manifest as vertical changes in the 1909-1936

leveling data. I cannot exclude the possibility that aseismic or weakly-seismic mobility of the structures

in the fold-belt prior to the earthquake sequence may have been responsible for all three events.

5.9 Conclusions

Precise leveling data and a fault model derived from detailed geological cross-sections, permit us

to calculate the slip distribution on the rupture surface of the 1931 Mach earthquake. I deduce that the

earthquake occurred on the 42 km wide (EW), 72 km long (NS) Dezghat/Bannh fault system west of Sibi.

The fault slipped in a reverse sense up to the east with maximum slip of 3.2 m and mean slip of 1.2 m.

Maximum slip coincides spatially with that predicted in elastic models driven by inferred interseismic

stresses, but themaximumslip is larger thanpredicted compared to themean slip of the fault. The leveling

data suggest that slip also occurred downdip of the rupture (approximately 1 m) either as afterslip or slip

in other events in the interval 1909–36.

A significant (3–5m) slip deficit remained near the interseismic locking line 4 years after the earth-

quake. This slip deficit may now, nearly 75 years later, have been reduced by aseismic processes sub-

sequent to the earthquake, or it may remain stored as elastic strain to drive future earthquakes. Two

mechanisms may act to prevent the accumulation of seismic deficit over multiple earthquake cycles. The

first is that slow, off-fault deformation, or pressure solution processes, act to reduce local stresses at the

locking line, and the second is that infrequent larger earthquakes mine an historically stored slip deficit,

accompanied by an incremental shift in the depth of the locking line. The first is testable in principle, in

that one could re-measure surviving points of the leveling line to determine whether the slip deficit re-

99

mains. In practice, current security issues in the region render this difficult. I favor, however, the second

mechanism: that stored elastic strain from one ormore previous earthquakesmay account for the 3.2m of

local slip observed updip from the inferred locking line in the 1931 earthquake. The local maximum slip is

easier to explain as an additional 1.5–2 m of slip inherited from strain un-released by former earthquakes,

than the alternative solution, that the slip in the earthquake should everywhere have been approximately

3 m in 1931 and that an approximately 2 m slip deficit remains on most of the fault.

In considering the details of theMach earthquake, I examined in numerical experiments the elastic

processes prevailing during the entire seismic cycle. I found that slip on a frontal thrust is always less than

the geodetic contraction rate of a fold and thrust belt, unless the frontal thrust consists of a planar fault.

In the case of a geometrically complex underlying thrust fault with variable dip, I find that partitioning

of slip to the frontal thrusts is reduced in proportion to the ratio of decollement-depth to shallow-thrust-

depth where these are separated by a ramp. This significant discrepancy between the geodetic loading

rate and the slip-potential of frontal faults of the fold belt is presumably responsible for thickening the

pile of sediments by folding and listric faulting. Partitioning in the Kirthar range, as elsewhere, results

in a significantly longer renewal time for earthquakes on the frontal fault of the range, than would be

derived from the geodetic convergence rate alone. Thus, although the weakly constrained 5 mm/yr GPS

convergence rate between Quetta and Sibi would result in a minimum renewal time for 1.2 m (average

slip) on a planar frontal thrust fault of 240 years, this study suggests that the presence of a decollement

at approximately 18 km depth would extend the recurrence interval for Mach-type earthquakes on the

Dezghat/Bannh fault system by a factor of 8, to approximately 2000 years.

Given that more than 90% of the seismic moment release in the region occurred between 1930 and

1935, I believe that sequential triggering of the three earthquakes occurred. The 1931 Sharigh earthquake

was clearly responsible for triggering the Mach earthquake 3 days later. The structural relationship be-

tween these two fault systems is discussed further in Chapter 6. In contrast, the “broadside” relation

between the Mach and Quetta rupture zones is consistent with an increase in Coulomb failure stress on

the Quetta fault at the time of the Mach earthquake. The 3.5 yr interval between the two earthquakes in-

dicates, however, that static stresses changes alone were insufficient to trigger the Quetta earthquake.

100

I hypothesize that the Mach earthquake reduced east-west stresses on the decollement/ramp system

that would have facilitated accelerated creep on the basal decollement beneath the Kirthar range. De-

formation between Mach and Quetta inferred from minor uplift and subsidence in the leveling data, are

consistent with strain changes accompanying decollement slip, although interpretation of these data are

non-unique.

Chapter 6

Bookshelf Faulting in the 2008 Ziarat Earthquake Sequence, Northern Baluchistan

6.1 Introduction

The Indian Plate is convergingwith the Eurasian Plate at a rate of 38mm/yr at the location ofHyder-

abad, India (Altamimi et al., 2007). Along the western boundary of the Indian Plate, this collision manifests

itself primarily as sinistral slip along the Chaman Fault System. Recent investigations of seismicity along

the Chaman Fault System in Baluchistan suggest that deformation is partitioned between sinistral motion

and range normal convergence (Chapters 5 and 4). In addition, recent geodetic estimates suggest that

overall sinistral rates are near the lower bound of geologic slip estimates (19.5 mm/yr, Chapter 4).

Some of this slip partitioning along the plate boundary manifests itself as diffuse deformation in

the Sulaiman Lobe, a southward verging salient produced by the northward translation of the semi-rigid

Katawaz block by the Chaman Fault System (Figure 6.1; Haq and Davis, 1997; Bernard et al., 2000). The south-

ward extrusion of the Sulaiman Lobe is accommodated along its eastern margin by the left-lateral Kingri

fault and related structures (Figure 6.1; Rowlands, 1978). Similarly, seismicity in the Quetta Syntaxis, along

the western margin of the Sulaiman Lobe, suggests the presence of a dextral feature analogous to the

Kingri fault, although there are no mapped faults with dextral offset in this region (Banks and Warburton,

1986; Bannert et al., 1992; Schelling, 1999b).

Seismicity in the Quetta Syntaxis occurs in a NW-SE oriented band approximately 25 km wide and

stretching 100 km from Pishin in the NW to near Harnai in the SE (Figure 6.2 and Table 6.1). Near the

town of Pishin, the southern end of the seismically quiet Katawaz Block marks one end of the seismic

zone, while to the SE, seismicity becomes more diffuse and focal mechanisms become dominantly thrust

102

66˚ 68˚ 70˚28˚

30˚

32˚

Kat

awaz

Blo

ck

Gh

aza

ban

d F

au

ltC

ham

an

Fau

lt

Kingri Fault

Bannh Fault

Dezghat Fault

Kir

thar

Ran

ge

SulaimanLobe

QuettaSyntaxis

Su

laim

an

Ran

ge

29 mm/yr

Figure 6.1: Map of the Sulaiman Lobe andnorthernKirthar Range of Pakistan, highlighting themajor faultsof region. The Bannh and Dezghat faults last ruptured during the 1931 Mach earthquake. The GhazabandFault is presumed to have last ruptured during the 1935 Quetta Earthquake and the Chaman Fault last rup-tured in 1892 and 1976. The Katawaz Block of Haq and Davis (1997) is outlined with a dashed line. The threestars indicate the locations of the twomainshocks and the largest aftershock of the 2008 Pishin Earthquakesequence. The Kingri Fault is a sinistral fault and is presumed to enable the southward extrusion of theSulaiman Lobe (Rowlands, 1978).

103

faulting (Figure 6.3; Bernard et al., 2000). At least 4 earthquakes larger than Mw 6 have occurred in this

seismic belt in the past century (Table 6.1), the 1931 Mw 6.8 Sharigh earthquake (Chapter 5), the 1997

Mw 7.1Harnai earthquake (Khan, 1998; Bernard et al., 2000) and both 2008Mw 6.4 Ziarat earthquakes (this

Chapter). In addition, numerous moderate earthquakes (Mw > 5) have occurred in the region over the

past century (Figures 6.2 and 6.3). There is some indication that the epicenter of the 1909M ≈ 7 Kachhi

earthquake was close to the location of the 1997Mw 7.1 Harnai earthquake (Engdahl and Villasenor, 2002).

Table 6.1: Historical earthquakes in the Quetta Syntaxis. An additional 5 earthquakes with magnitudesbetweenMw 5.1 andMw 5.4 occurred during the Oct.–Dec. 2008 aftershock sequence but are unlisted.

Epicenter MagnitudeNumber Date Epicenter MagnitudeSource Source1 20 Oct. 1909 68.0E, 30.0N ISC 7.2 ISC2 24 Aug. 1931 67.7E, 30.2N ISC 6.8 ISC3 29 Sep. 1941 67.2E, 30.7N ISC 5.4 ISC4 16 Jun. 1976 67.2E, 30.7N ISC 5.1 ISC6 16 Nov. 1993 67.0891E, 30.8024N This Study 5.6 This Study7 27 Feb. 1997 67.9875E, 29.9932N This Study 7.2 ISC8 28 Oct. 2008 67.3825E, 30.5012N This Study 6.4 This Study9 29 Oct. 2008 67.5297E, 30.4659N This Study 6.4 This Study10 9 Dec. 2008 67.4416E, 30.4024N This Study 5.7 This Study

The identification of rupture planes for earthquakes in the Quetta Syntaxis has been hampered by

the mismatch between the expected dextral sense of deformation, the actual trend of mapped surface

structures and the ambiguity of fault planes from seismically determined moment tensors. Focal mecha-

nisms throughout theNW-SE trending seismic bandnorth of Quetta showdominantlyN-S trending P-axes,

consistent with the current day northward motion of the Indian plate relative to Eurasia (Figure 6.3; Al-

tamimi et al., 2007)). As one moves from NW to SE along this seismicity band, there is a continuous change

in mechanism from dominantly strike-slip in the NW to range-normal thrusting in the SE where mapped

fault orientations becomemore favorable to thrust fault activation (Bernard et al., 2000). Given the seismic

productivity of this region, little work has been done to understand how shear in the Quetta Syntaxis is

accommodated tectonically.

104

66˚ 67˚ 68˚

30˚

31˚

Quetta

Ziarat

Sharig

Mach

Sibi

Harnai

Chaman

Pishin

1

3

2

4

5

6

78

9

Figure 6.2: Map of the Sulaiman Lobe and northern Kirthar Range of Paksitan, showing the location oftowns mentioned in the text. Historical earthquakes in the Quetta Seismic zone, 1900–2010. Numbersreference dates, epicenters and magnitude listed in Table 6.1. Only the three largest earthquakes fromthe 2008 aftershock sequence are shown on the map.

105

66˚ 68˚ 70˚28˚

30˚

32˚

20±2 mm/yr

Figure 6.3: Spatially averaged GPS velocities with respect to the stable Indian Plate and centroid momenttensors from the Global CMT (Dziewonski et al., 1981) withMw > 5 since 1976. Filled regions are compres-sional quadrants of the best-fitting double couple. Note the lack of seismicity within the boundaries of theKatawaz Block (see Figure 6.1 for place names). Velocities are calculated as the weighted spatial averageof all regional GPS velocities within a 30’ grid. The location of each velocity average is calculated as themean of the locations within each grid.

106

In this study, I focus on an earthquake sequence beginning in late Oct. 2008. At 2309 UTC on 28 Oct.

2008, an Mw 6.4 earthquake struck the region 40 km NE of Quetta. This earthquake was followed by a

similarly sized earthquake 11 hours later, 15 km SE of the first shock. The aftershock sequence from these

two earthquakes consisted of more than 50 earthquakes larger thanMw 4 with a total additional seismic

moment equivalent to Mw 6.0 and lasted until mid-December 2008. In the following account, the two

principal earthquakes in the sequence will both be referred to as mainshocks due to their approximately

equal magnitudes. When necessary, I will distinguish them by their date. I begin by analyzing teleseismic

body-wave data for the two mainshocks along with the largest aftershock. Because of the complexity of

the interferometric data spanning both mainshocks, I proceed by inverting interferometric data for the

Mw 5.7 aftershock of 9 Dec. 2008. I will then utilize the precise location information for this aftershock

provided by InSAR inversion to revise the double-difference relocations for the two mainshocks. These

revised mainshock locations along with the fault planes obtained from teleseismic body-wave inversion

are then used as starting parameters for an inversion of the complex coseismic mainshock interferogram.

I infer that this earthquake sequence involved sinistral faulting along NE-SW trending faults. This

interpretation, combined with the results of an inversion of InSAR data for anMw 5.6, 40 km to the NW of

the 2008 earthquake sequence, suggests that overall dextral shear in the Quetta Syntaxis is accommodated

along en-echelon NE-SW trending sinistral faults through “bookshelf faulting”.

6.2 Tectonic Overview

The lobate shape of the fold-and-thrust belts of western Pakistan have been a source of intrigue for

structural geologists for at least the past half-century (Jones, 1961; Rowlands, 1978; Quittmeyer et al., 1984;

Banks and Warburton, 1986; Humayon et al., 1991; Jadoon et al., 1993, 1994; Jadoon and Kurshid, 1996; Haq and

Davis, 1997; Bernard et al., 2000). Analogue and viscoelastic modeling of the margin demonstrates that the

lobate structure as well as the strike of both the Kirthar and Sulaiman Ranges can be attributed to the

translation of a semi-rigid block (the Katawaz block) northward along the eastern edge of the Chaman

Fault System (Haq and Davis, 1997; Bernard et al., 2000). Southward extrusion of the Sulaiman Lobe is ac-

commodated by simple strike-slip faulting on its eastern margin along faults such as the Kingri Fault (Fig-

107

ure 6.3; Rowlands, 1978). Along the western margin of the Sulaiman Lobe, the convergence velocities and

directions between the Sulaiman Lobe and the northern Kirthar Range suggest that dextral shear accom-

modates their differential shortening rates (Figure 6.3).

The Quetta Syntaxis is located at the transition between the fold-and-thrust belts of theN-S striking

Kirthar Range and the southward verging Sulaiman Lobe. The N-S strike of both the Kirthar Range and

the Sulaiman Range are nearly parallel to the northward velocity of the Indian Plate relative to Eurasia

(29 mm/yr at N7E, Figure 6.1; Altamimi et al., 2007). Proceeding east from the Kirthar Range, the trend of

mapped structures rotates to a nearly NW-SE azimuth at the apex of the Quetta Syntaxis and becomes

perpendicular to the India-Eurasia convergence direction before rotating back to a N-S orientation in the

Sulaiman Range (Figure 6.1). The intersection between the N-S trending Kirthar Range and the NW-SE

trending structures of the western Sulaiman Lobe, at the apex of the Quetta Syntaxis, coincides with the

most seismically active region of the western Indian Plate margin.

6.3 Data and Methods

6.3.1 Double-difference Relocations

I relocate both mainshocks of the 2008 Ziarat earthquake sequence along with its aftershock se-

quence through Jan. 2009 using P-wave phase data published in the monthly National Earthquake Infor-

mation Center Earthquake Data Records and a double-difference methodology (Waldhauser and Ellsworth,

2000). To increase the robustness of the relocations, I also include phase data from the nearby 1997Mw 7.1

Harnai earthquake along with its aftershock sequence through Apr. 1997.

6.3.2 Teleseismic Body-wave Modeling

Using IRIS’s WILBUR II (http://www.iris.edu/wilber), I obtained broad-band seismic data from

Global Seismic Network stations in the distance range 30–90◦ to avoid complexities from mantle triplica-

tions and the outer core. I then selected the best P and SHwaveformswith an emphasis on good azimuthal

coverage. Waveforms were low-passed filtered using a 3-pole Butterworth filter with a corner frequency

108

of 0.08 Hz and down-sampled to 0.5 s sample spacing. I then use the programMT5 (McCaffrey et al., 2000) to

simultaneously invert P and SHwaveforms for the best fitting double couple point source using aweighted

least-squares approach. I use a simple half-spacemodel for crustal structurewithwave speeds ofVP = 6.5

km/s and VS = 3.7 km/s and a density of ρ = 2.8 kg/m3. I use a mantle attenuation parameter (t∗) value

of 1 s for P-waves and 4 s for SH-waves.

6.3.3 InSAR

To study deformation related to the 2008 Ziarat earthquake sequence, I utilize 6 ascending pass SAR

scenes spanningNov. 2007–Jan. 2009 from the European Space Agency’s Envisat satellite (Figure 6.4), while

deformation from the 16 Nov. 1993 Pishin earthquake, was captured using 2 descending pass SAR scenes

spanning Sep. 1993–Nov. 1993 from the European Space Agency’s ERS-1 satellite (perpendicular baseline of

-22 m). Descending pass Envisat data covering the 2008 Ziarat earthquake sequence could not be used due

to temporal decorrelation over the coseismic region. A 90 m resolution Digital Elevation Model was con-

structed from Shuttle Radar Topography Mission version-2 data (Farr et al., 2007) to remove topographic

fringes. Interferograms were produced using the ROI PAC InSAR software package developed at the Jet

Propulsion Laboratory in Pasadena, CA (Rosen et al., 2004). Interferograms were sampled with 4 looks in

range and 20 looks in azimuth to produce 80 m × 80 m resolution cells, filtered using a power spectral

method (Goldstein and Werner, 1998) and unwrapped using a least squares methodology. Unwrapped in-

terferograms were then subsampled using a resolution-based methodology (Lohman and Simons, 2005) and

inverted for fault location and orientation using Powell’s conjugate gradient descent method with Monte

Carlo restarts (Powell, 1964; Clarke et al., 1997; Funning et al., 2007).

6.3.4 GPS Data

Continuous GPS measurements in the region of the Ziarat earthquake began in 2005 and have been

supplemented by campaign GPS measurements since 2007. Campaign measurements from 13 sites with

locations predominantly north of Quetta have been measured at least twice in the period 2007–2010 and

compared to continuous measurements made in Karachi (KCHI), Sukkur (SIBA), Peshawar (NCEG) and

109

−200

0

200

400

Pe

rpe

nd

icu

lar

Ba

se

line

(m

)

January April July October January

2008Date

Envisat Track 213 Frames 585−621

2 M

w 6

.4 e

arth

quak

es

Mw

5.7

ear

thquak

e

Figure 6.4: Scene acquisition date versus perpendicular baseline for Envisat track 213 frames 585–621.Circles represent Envisat Image Mode 6 SAR scenes while lines represent SAR interferograms. Scenesdenoted by gray circles are heavily contaminated with topographically correlated atmospheric signalsand were not used. Solid black lines denote interferograms used to invert for fault parameters. Solid graylines denote coseismic interferograms which were not used. Vertical dashed lines mark the times of the28–29 Oct. 2008 mainshocks and 9 Dec. 2008 aftershock discussed in the text.

110

Quetta (QTAG and QTIT) (Figure 6.5). Continuous GPS stations in Pakistan are operated from flat-roofed

concrete frame buildings while campaign points are measured on bipods set on stainless steel screws ce-

mented into exposed rock. GPS observations were recorded using Trimble NetRS, 5700 and R7 receivers

using a 30 s sampling rate, and processed using an elevation cutoff angle of 10◦. Campaign data have du-

rations of 3–7 days from each site for each occupation. The daily data from these sites were processed

along with data from 4 continuous stations in Pakistan and 10 regional IGS stations using GAMIT version

10.35 (King and Bock, 2002). The regional solutions were then combined with global solutions from SOPAC

(http://sopac.ucsd.edu) using GLOBK/GLORG version 5.17 (Herring, 2002) to determine time series and

velocities consistent with the ITRF2005 reference frame. These velocities were then transformed into an

Indian plate-fixed reference frame using pole of rotation parameters determined by Altamimi et al. (2007).

6.3.5 Macroseismic Observations

For some of the events in the Quetta Syntaxis from the early 20th century, specifically the 1931

Sharigh earthquake, macroseismic data provide the best constraint on epicentral location. In order to

assist in the interpretation of these historical earthquakes, I analyze the macroseismic data available for

the 28 and 29 Oct. 2008 earthquakes and compare estimates of their epicentral locations with the revised

double-difference epicenters (Figure 6.6). Martin and Szeliga (2010) provides uniformly assessed intensities

for both the 28 and 29 Oct. earthquakes, while Ambraseys and Douglas (2004) provides intensities for the

1931 Sharigh earthquake. Estimates of epicentral location are calculated using the methodology outlined

in Chapter 3.

6.4 Interpretational Procedure

The apparent NW-SE alignment of the epicentral locations for the two mainshocks coupled with

their shared NW-SE trending nodal planes suggests that both mainshocks represent sequential rupture of

a single NW-SE trending fault. While such an interpretation is consistent with the observation of dextral

shear in the southern Sulaiman Lobe (Figure 6.3), no surface faults with this trend are mapped in the

111

67˚ 68˚ 69˚

30˚

31˚

LORI

QLAS

SANJ

CHTR

HRNIHRNI

SHRGSHRG

ZART

MUSB

SURB

QTIT

KHAL

LAKP

QILA

SHBG

KACH

KHST

GULH

QTAG

SARN

10 mm/yr

A

30˚

31˚

30 mm

28−29 Oct. 2008 Coseismic

B

67˚ 68˚ 69˚

30˚

31˚

30 mm

9 Dec. 2008 Coseismic

D

30 mm

28−29 Oct. 2008 Residual

C

67˚ 68˚ 69˚

30 mm

9 Dec. 2008 Residual

E

Figure 6.5: Interseismic velocities, coseismic offsets and residuals for the 28–29 Oct. 2008 earthquakes andthe 9 Dec. 2008 earthquake. A.) Interseismic velocities relative to the stable Indian Plate. Thick black lineswithout arrows represent regional faults (see Figure 6.1). B.) Coseismic offsets and C.) residuals from the28–29 Oct. 2008 earthquake. Displacements for stations KHST and SHRG are poorly defined due to lownumber of post-seismic observations. Stations ZART and CHTR were established in 2009 and thereforehave no pre-seismic position measurements. Black lines represent the rupture planes determined frominversion of InSAR data. D.) Coseismic displacements and E.) residuals for the 9 Dec. 2008 earthquake.The proximity of station KACH to the epicenter combined with fortunate post-seismic occupation timingmakes this the only station for which I am able to estimate displacements. Black lines represent rupturedetermined from inversion of InSAR data. The error ellipses represent formal uncertainties for the co-seismic displacements as measured from the time series for each station and certainly represent a bestcase scenario. The residual displacements are calculated by removing the best-fitting coseismic modeldetermined from inversion of InSAR data.

112

66˚ 67˚ 68˚ 69˚

28˚

29˚

30˚

31˚

A

66˚ 67˚ 68˚ 69˚

B

66˚ 67˚ 68˚ 69˚

C

50 km

Figure 6.6: Epicentral locations for the 24 Aug. 1931 Sharigh earthquake and the 28 and 29 Oct. 2008 Ziaratearthquakes determined from shaking intensity data. Locations are determined using the methodologyoutlined in Chapter 3. The contours represent the 50%and 67%confidence contours for epicentral locationcalculated using parameters listed in Bakun (1999). In each subfigure, filled circles indicate the locations offelt reports, the star indicates the instrumentally determined epicenter and the center of the innermostcontour represents the preferred macroseismic estimate of epicenter. Intensity data are from Martin andSzeliga (2010). A.) Epicenter of the 24 Aug. 1931 Sharigh earthquake as determined from macroseismicdata. B.) Epicenter of the 28 Oct. 2008 earthquake as determined frommacroseismic data. C.) Epicenter ofthe 29 Oct. 2008 earthquake as determined from macroseismic data.

113

region (Banks andWarburton, 1986; Bannert et al., 1992; Schelling, 1999b). Mapped surface faults in the Quetta

Syntaxis consist primarily of E-W trending thrust sheets, however the dip of NW-SE trending nodal planes

of the moment tensors (Figure 6.7) is inconsistent with reactivation of these thrust sheets in a dextral

sense.

16 Nov. 1993

28 Oct. 2008

29 Oct. 2008

9 Dec. 2008

Body−wave CMT InSAR

Decreasing Frequency

Figure 6.7: Graphical comparision of moment tensor solutions from inversion of teleseismic body-wavedata, the Global CMT (Dziewonski et al., 1981) and inversion of InSAR data. Each inversion method is sensi-tive to deformation in different frequency bands. To illustrate this, moment tensors are arranged, fromleft to right, in order of sensitivity to decreasing frequencies (increasing periods) of radiated energy. Incases where more than one subevent is inverted for, the moment tensor for the subevent with the largestcontribution to the total moment is shown.

While in general there is an ambiguity between the two nodal planes of a double couple focal mech-

anism determined from teleseismicwaveformdata alone, formost earthquakes large enough to be imaged

by interferometric SAR techniques, the finiteness of the rupture plane coupled with the look angle of the

radar produces a deformation pattern that often allows for the determination of the actual rupture plane.

I begin by determining the best-fitting fault plane solutions using teleseismic body-wave inversion.

These fault planes are then used as starting parameters in the inversion of the InSAR data. I first exam-

ine the 2 Dec. 2008–6 Jan. 2009 interferogram (Figure 6.8), inverting for deformation from the 9 Dec. 2008

aftershock, and use its fault location as ground-truth to translate the double-difference hypocentral relo-

cations for both mainshocks. These revised mainshock locations, coupled with fault plane solutions from

teleseismic body-wave inversion of each earthquake are then used to guide the inversion of the complex

mainshock interferogram (Figure 6.9).

114

67˚00' 67˚30'

30˚00'

30˚30'

A B

67˚00' 67˚30'

C

0 π 2π

Figure 6.8: Envisat interferogram of scenes from 2 Dec. 2008 and 6 Jan. 2009. One fringe corresponds to28 mm of change in range. Solid arrow indicates the flight direction of the satellite and outlined arrowdenotes the look direction of the satellite. A.) Original interferogram. B.) Preferred coseismic elastic dislo-cation model. C.) Interferogramwith coseismic model removed. Black line denotes the surface projectionof the up-dip edge of the fault identified from inversion of A.

67˚00' 67˚30'

30˚00'

30˚30'

A B

67˚00' 67˚30'

C

0 π 2π

Figure 6.9: Envisat interferogram of scenes from 6 May 2008 and 2 Dec. 2008. One fringe corresponds to28 mm of change in range. Solid arrow indicates the flight direction of the satellite and outlined arrowdenotes the look direction of the satellite. A.) Original interferogram. B.) Preferred coseismic elastic dislo-cationmodel. C.) Interferogramwith coseismicmodel removed. Black lines denotes the surface projectionof the up-dip edge of the fault identified from inversion of A.

115

6.4.1 The 9 Dec. 2008 Aftershock

The largest aftershock of the sequence occurred at 22:52:37 GMT on 9 Dec. 2008. Waveform fits

to this aftershock are straightforward and show primarily strike-slip on high angled fault planes (Figure

6.10). Two subevents separated by at most 12 s are suggested by the data. While the first subevent is

responsible for the greatest moment release, the second subevent is not insignificant, and is responsi-

ble for an additional moment release equivalent to 46% of the first subevent. The Global CMT solution

(Dziewonski et al., 1981) indicates a steeply dipping plane, similar to the first subevent of the body-wave

inversion, and has a total moment nearly equivalent to the sum of both subevents from the body-wave

inversion (Table 6.2). Fault plane parameters as determined by the Global CMT project (Dziewonski et al.,

1981, http://www.globalcmt.org), teleseismic body-wave inversion and InSAR inversion are shown in

Table 6.2.

Table 6.2: Comparison of the fault plane parameters for the preferred double-couple rupture plane of the9 Dec. 2008 aftershock. Strike, dip, rake and depth are for the preferred fault plane from the double-couplewith the largest contribution to the total moment. Moment is the total moment of the entire event. Eachinversionmethod is sensitive to deformation in different frequency bands. To illustrate this, solutions arearranged vertically from shortest (Body-wave) to longest (InSAR) period of sensitivity to radiated energy.For a visual comparison of each solution, see Figure 6.7.

Strike Dip Rake Depth Moment(deg.) (deg.) (deg.) (km) (N-m)

Body-wave 57 80 -14 7 3.8×1017

CMT 62 75 0 15.9 3.98×1017

InSAR 241.8 89.5 17.1 2.6 5.58×1017

Examination of perpendicular baseline versus time for scenes from Envisat ascending track 213

frame 585 (Figure 6.4) reveals only one possible scene combination displaying deformation solely from

the 9 Dec. 2008 earthquake. Using the fault plane results of the teleseismic body-wave inversion and the

epicenter from the double-difference relocation as starting parameters for inversion of this InSAR data,

I find a best fitting rupture plane that trends NE-SW. While the number of residual fringes in the 2 Dec.

2008–9 Jan. 2009 interferogram is low (< 1 fringe, Figure 6.8), suggesting that nearly all of the coseismic

deformation has been modeled, there remain numerous fringes away from the coseismic region likely

116

0 60s

0 14s STF

A

TATO 

A

B

NACB 

B

C CHTO 

CD

KKM 

D

E

SBM 

E

F

WR6 

F

G

WR5 

G

H

WR4 

H

I

WC2 

I

J

WR3 

J

K

WRAB 

K

L

WC1 

L

M

KUM 

M

N

KOM 

N

O

PSI 

O

P

KMBO 

P

Q

TSUM 

Q

RMBAR 

RS

FURI 

ST

IDI 

T

U

TIP 

U

V

TIR 

V

A

YAK 

A

B

ULN 

B

C

YSS 

C

D

ASAJ 

D

E

NACB 

E

FSSLB 

FG

LBTB 

G

H

KMBO 

H

I

TSUM 

I

J

MBAR 

J

KFURI 

K LDBIC 

L

M

DIVS 

M

9 Dec. 20081:57/80/346/7/2.614e172:185/62/22/10/1.193e17

Figure 6.10: Lower hemisphere projection of the moment tensors from the inversion of teleseismic body-waves for the 9 Dec. 2008 aftershock. Fault plane information for each subevent are listed in the header asevent number, strike, dip, and rake in degrees, depth in km and moment in N-m. Seismic station namesare printed vertically and to the left of each waveform. Seismic station locations on the focal sphere aredenoted by upper-case letters and correspond to the letter indicated between the station name and thewaveform trace. Upper plot shows P-wave focal sphere and waveforms, while the lower plot shows SH-wave focal sphere and waveforms. Amplitudes have been normalized to highlight the agreement betweenthe data (solid line) and the synthetic waveforms (dashed line). The source-time function along with thetime scale for each waveform is shown beneath the P-wave data for station KMBO.

117

due to atmospheric turbulence (Massonnet and Feigl, 1995). As confirmation of the InSAR derived coseismic

model, I amable to predict nearly 90%of thehorizontal displacement observed at GPS stationKACH (Figure

6.5).

A visual comparison of all threemoment tensors is presented in Figure 6.7 and suggests good agree-

ment. The disagreement between the depth estimated from the body-wave inversion and the InSAR inver-

sion suggests a difference between the location of rupture initiation and the location of greatest moment

release.

While the results of the double-difference relocations are self-consistent, they contain a systematic

bias that results in a translation of the entiremainshock-aftershock cluster relative to its true location. To

estimate this translation, I utilize the center of the surface projection of the InSAR derived fault plane for

the 9 Dec. aftershock as a ground-truthmeasurement. I calculate the offset between the double-difference

location and the InSAR derived fault center as approximately 15 km at an azimuth of 293◦. I then apply

this translation to the entire mainshock-aftershock cluster to derive a revised set of epicentral locations

(Figure 6.11). From a purely geometric perspective, choosing the center of the surface projection of the

InSAR derived fault leads to a roughly 6.5 km uncertainty in relocation parallel to the fault plane (NE-SW

direction), and 1.2 km perpendicular to the fault plane (NW-SE direction).

6.4.2 28–29 Oct. Mainshocks

The first mainshock occurred at 23:09:57 GMT on 28 Oct. 2008, and was preceded 36minutes earlier

by an Mw 5.2 foreshock. Waveform fits to this mainshock are well described using a single large event

with a source-time duration of 8 s followed closely by a smaller event (∼ 20% smaller) with a simple

triangular source-time function and a duration of 2 s (Figure 6.12). Both subevents have similar fault

planes, suggesting the point source assumption is not entirely valid and instead, slip occurred over a fault

plane of finite extent. The preferred rupture plane is listed in Table 6.3.

In contrast to the first mainshock, waveforms for the second mainshock, occurring 11 hours later,

are more complex. The best-fitting moment tensor for the 29 Oct. 2008 earthquake requires at least two

subevents (Figure 6.13). Forward modeling of seismic waveform data from two stations at regional dis-

118

66˚ 68˚

30˚

Quetta

Ziarat

Sharig

Mach

Sibi

Harnai

Chaman

Pishin

Figure 6.11: Revised double-difference earthquake relocations for all events in the region during the peri-ods Feb. 1997–Mar. 1997 and Oct. 2008–Jan. 2009. Earthquakes during this time period were relocated us-ing phase data from the USGSmonthly PDE using the double differencemethod ofWaldhauser and Ellsworth(2000). Double difference locations for the 9 Dec. 2008 Mw 5.7 aftershock were then compared with thelocation derived from inversion of the interferogram in Figure 6.8 to obtain a shift parameter. Reviseddouble difference epicenters were then obtained by applying this shift parameter to all of the double dif-ferenced earthquakes.

Table 6.3: Comparison of fault plane parameters for the preferred double couple rupture plane of the 28Oct. 2008 mainshock. Strike, dip, rake and depth are for the fault plane with the largest moment release.Moment is the total moment of the entire event. Solutions are arranged vertically from shortest (Body-wave) to longest (InSAR) period. For a visual comparison of each solution, see Figure 6.7.

Strike Dip Rake Depth Moment(deg.) (deg.) (deg.) (km) (N-m)

Body-wave 25 58 36 8 4.6×1018

CMT 37 81 18 17.2 5.08×1018

InSAR 45.9 38.8 54.7 11.5 2.5×1018

119

0 60s

0 14s STF

A

SPBG 

A

B

TLY 

B

C

HIA 

C

D

INCN 

D

E

SSE 

E

F

ENH 

F

GNACB 

GHTPUB 

HI

KMI 

I

J

PMG 

J

K

SBM 

K

L

IPM 

L

M

DGAR 

M

N

RER 

N

O

MSEY 

O

P

BOSA 

P

Q

LBTB 

Q

R

MBAR 

R

S

FURI 

S

TMSKU 

T UDBIC 

UV

WDD 

V

W

CEL 

W

X

VSL 

X

Y

PAB 

Y

Z

TUE 

Z

[

GRFO 

[

\

KBS 

\

A

EYAK 

A

B

YAK 

B

C

MDJ 

C

D

ERM 

D

E COEN 

EF

KSM 

F

G

WRAB 

G

H

IPM 

H

I

FITZ 

I

J

DGAR 

J

K

FOMA 

K

L

ABPO 

L

M

BOSA 

M

N

SUR 

N

O

TSUM 

O

P

MBAR 

P

Q

VSL 

Q

R

PVAQ 

R

S

VTS 

S

T

TIRR 

T

U

SSB 

U

V

BFO 

V

28 Oct 20081:25/58/36/8/3.754e182:28/56/31/9/8.514e17

Figure 6.12: Lower hemisphere projection of the moment tensors from the inversion of teleseismic body-waves for the 28 Oct. 2008 aftershock. Fault plane information for each subevent are listed in the headeras event number, strike, dip, and rake in degrees, depth in km andmoment in N-m. Seismic station namesare printed vertically and to the left of each waveform. Seismic station locations on the focal sphere aredenoted by upper-case letters and correspond to the letter indicated between the station name and thewaveform trace. Upper plot shows P-wave focal sphere and waveforms, while the lower plot shows SH-wave focal sphere and waveforms. Amplitudes have been normalized to highlight the agreement betweenthe data (solid line) and the synthetic waveforms (dashed line). The source-time function along with thetime scale for each waveform is shown beneath the P-wave data for station RER.

120

tances (KBL and NIL) proved difficult and the resulting best-fit mechanism differed considerably from

both the Global CMT solution and the best-fit body-wave mechanism, further suggesting a complex rup-

ture sequence (S. Roecker, personal communication Dec. 2009). Analysis of the rupture history for this

event is ambiguous due to the non-uniqueness inherent in the placement and depth determination for

each subevent. The preferred double couple rupture plane is listed in Table 6.4.

Table 6.4: Comparison of fault plane parameters for the preferred rupture plane of the 29 Oct. 2008 main-shock. Strike, dip, rake and depth are for the fault plane with the largest moment release. Moment is thetotal moment of the entire event. Solutions are arranged vertically from shortest (Body-wave) to longest(InSAR) period. For a visual comparison of each solution, see Figure 6.7.

Strike Dip Rake Depth Moment(deg.) (deg.) (deg.) (km) (N-m)

Body-wave 228 75 -13 9 4.1×1018

CMT 233 88 -22 12 5.39×1018

InSAR 214 40 -19.9 7.5 1.48×1018

Three of the radar acquisitions during the Summer of 2008 (July, August and September 2008) con-

tain strong atmospheric signals near to and south of the coseismic rupture area. These atmospheric signals

were large enough to preclude their use in forming interferograms suitable for modeling. Of the remain-

ing four radar acquisitions, it is possible to form only one interferogram containing deformation solely

from the 28–29 Oct. earthquakes and only one interferogram containing deformation solely from the 9

Dec. 2008 aftershock (Figures 6.4, 6.8, and 6.9).

In order to invert the complex interferogram containing the deformation signal from the twomain-

shocks (Figure 6.9), I utilize the epicentral location provided by our revised double-difference relocation

scheme as well as fault parameters derived from teleseismic body-wave inversion (Tables 6.3 and 6.4).

The number of residual fringes is large in the 6May 2008–2 Dec. 2008 interferogram, reaching up to

2 full fringes in the region near the epicenter of the 28 Oct. mainshock. In addition, slip modeling of the 29

Oct. mainshock predicts deformation to the west of the northern extent of faulting that is not seen in the

data. Given the complexity of the coseismic deformation field, I feel that little additional improvement is

possible.

121

0 60s

0 14s STF

A

INK 

A

B

MCK 

B

C

HIA 

C

D

BJT 

D

E

JNU 

E

F KMI 

FG

KOM 

G

H

COCO 

H

I

DGAR 

I

J

MSEY 

J

K

FOMA 

K

L

ABPO 

L

MTAM 

M

N

TIRR 

N

O

SCHQ 

O

P

KBS 

P

Q

ALE 

Q

A

KDAK 

A

B

ULN 

B

C

SSE 

C

D KMI 

DE PMG 

EF

COEN 

F

G

WB2 

G

H

MORW 

H

I

FOMA 

I

J

SUR 

J

K

FURI 

K

LDBIC 

LM

PAB 

M

N

VSU 

N

O

KBS 

O

P

ALE 

P

29 Oct 20081:228/75/347/9/2.676e18

2:215/81/5/11/1.42e18

Figure 6.13: Lower hemisphere projection of the moment tensors from the inversion of teleseismic body-waves for the 29 Oct. 2008 aftershock. Fault plane information for each subevent are listed in the headeras event number, strike, dip, and rake in degrees, depth in km andmoment in N-m. Seismic station namesare printed vertically and to the left of each waveform. Seismic station locations on the focal sphere aredenoted by upper-case letters and correspond to the letter indicated between the station name and thewaveform trace. Upper plot shows P-wave focal sphere and waveforms, while the lower plot shows SH-wave focal sphere and waveforms. Amplitudes have been normalized to highlight the agreement betweenthe data (solid line) and the synthetic waveforms (dashed line). The source-time function along with thetime scale for each waveform is shown beneath the P-wave data for station DGAR.

122

Additional information on the location of rupture planes andpossible causal faults is provided in the

form of offset fringes and localized decorrelation in the interferogram (Figure 6.9A). These offset fringes

parallel themapped trace of the Gundak Rud Fault (Schelling, 1999b) and field reconnaissance in this region

in the weeks following the mainshocks identified ground cracks and slumping along the mapped trace of

the Gundak Rud fault. Near 30.4525N, 67.2524E, the sense of motion for these surface offsets was primarily

vertical with a maximum offset of 62 cm and was consistent with topographic slumping. InSAR fringe off-

sets suggest that minor surface rupture extended for at least 22 km along the Gundak Rud fault. However,

security concerns prevented detailed field mapping along the remainder of the fault.

The location of the surface projection of the causal fault for the second mainshock as determined

from the InSAR data lies within and parallels the Kan Tangai (Figure 6.14). Field investigations in the

Kan Tangai in the weeks following the mainshocks identified numerous N-S oriented ground cracks but

revealed no obvious surface rupture. Landsliding in the Ghazji shales near the Kan Tangai region, along

with the nearly complete destruction of the villages in the Kach and Ziarat valleys resulted in extensive

decorrelation surrounding the epicentral region of the second mainshock (Figure 6.9).

Figure 6.14: Photograph of the rupture zone of the 29 Oct. 2008 Ziarat Valley earthquake courtesy of DinMohammad Kakar. View looking south into the Kan Tangai (Stone Gorge) from the village of Wam. Nosurface rupturewas observed in the gorge, but numerous N-S trending surface crackswere apparent alongthe roads and hill-slopes.

123

The results of the inversion of the InSAR data are shown in Figure 6.9C. While residual fringes re-

main, I have modeled the regions with the greatest fringe gradient surrounding the surface offsets along

the Gundak Rud as well as fringes south of the decorrelated region south of the Kan Tangai. North of

the Kan Tangai, the residual interferogram shows additional fringes resulting frommismodeling of the 29

Oct. earthquake suggesting that uniform slip on a single fault is a poor approximation for this complex

earthquake. Removal of the coseismic displacement field determined from InSAR data from our GPS ob-

servations yields a 71% reduction in the coseismic signal at GPS station KACH but only a 25% reduction at

SURB (Figure 6.5).

6.4.3 16 Nov. 1993 Earthquake

The only previous earthquake to occur in the Quetta Syntaxis shear zone for which InSAR data

are available is a moderate magnitude earthquake that struck the region immediately north of Pishin

at 15:52:48 GMT on 16 Nov. 1993. The epicenter of this earthquake lies very close to the southern edge

of the Katawaz Block (Figure 6.2). Since seismicity becomes sparse to the NW of this region, I consider

this earthquake to represent the northwestern edge of the Quetta Syntaxis shear zone. Inversion of a 2.5

month interferogram covering the epicentral region of this earthquake indicates rupture along a NE-SW

trending, steeply dipping sinistral fault (Figure 6.15 and Table 6.5).

Table 6.5: Comparison of fault plane parameters for the preferred rupture plane of the 16 Nov. 1993 earth-quake. Strike, dip, rake and depth are for the fault plane with the largest moment release. Moment isthe total moment of the entire event. Each inversion method is sensitive to deformation in different fre-quency bands. To illustrate this, solutions are arranged vertically from shortest (CMT) to longest (InSAR)period of sensitivity to radiated energy. For a visual comparison of each solution, see Figure 6.7.

Strike Dip Rake Depth Moment(deg.) (deg.) (deg.) (km) (N-m)

CMT 24 89 -13 33.0 3.15×1017

InSAR 204.42 83.2 3.4 2.25 2.57×1017

Of further interest is the additional capture of a small earthquake located in the town of Pishin

south of the 16 Nov. rupture and centered along the northernmost trace of the Ghazaband fault. Only one

124

67˚00' 67˚30'

30˚30'

31˚00'

67˚00' 67˚30'

0 π 2π

Figure 6.15: ERS-1 interferogram of scenes from 13 Sep. 1993 and 22 Nov. 1993. One fringe corresponds to28mm of change in range. A.) Original interferogram. Black circle indicates the location of anMb 4.2 thatoccurred on 30 Oct. 1993 along the northernmost trace of the Ghazaband Fault. Solid arrow indicates theflight direction of the satellite and outlined arrow denotes the look direction of the satellite. B.) Preferredcoseismic elastic dislocation model. C.) Interferogram with coseismic model removed. Black line denotesthe surface projection of the up-dip edge of the fault identified from inversion of A.

125

additional earthquake is listed in the ISC catalog during the timespan of the interferogram, an Mb 4.3

earthquake on 30 Oct. 1993, suggesting that it is the source of the observed deformation.

6.5 Discussion

6.5.1 Historical Seismicity and Shear Zone Extent

In order to estimate the dimensions of the Quetta Syntaxis shear zone, one must consider the dis-

tribution and similarity of focal mechanisms within the shear zone. Data from the Global CMT (Dziewonski

et al., 1981) show that the 1997 Harnai earthquake and its largest aftershocks are primarily N-S thrust

faulting on shallowly dipping planes (Figure 6.3; Khan, 1998; Bernard et al., 2000). These mechanisms are

inconsistent with dextral shearing, indicating that the northwestern extent of the aftershock region of

the 1997 Harnai earthquake (Figure 6.11) is the likely terminus of the Quetta Syntaxis shear zone.

Examination of Figure 6.11 also reveals a gap in seismicity immediately northwest of the 1997 Har-

nai earthquake aftershock zone. This seismic gap extends for 30 km and separates the 1997 Harnai earth-

quake sequence from the 2008 Ziarat earthquake sequence. This gap, centered on the town of Sharigh,

also coincides with the most likely location for the 1931 Sharigh earthquake (Table 6.1 and Figure 6.6A).

As discussed in Chapter 5, the 1931 Sharigh earthquake is considered to be a foreshock to the 1931

Mach earthquake, whose epicenter is located immediately to the south of the Quetta Syntaxis Shear Zone,

in the Bolan River Valley. The 66 hour temporal spacing between theMw 6.8 Sharigh earthquake and the

Mw 7.1 Mach earthquake strongly suggests triggering, possibly due to Coulomb stress transfer (Chapter

5; Stein et al. (1994)). Comparisons between Coulomb stress transfer from an E-W oriented thrust faulting

mechanism and a NE-SW oriented sinistral faulting mechanism along inferred faults near the town of

Sharigh (Figures 6.1 and 6.16) show that a thrust faulting mechanism transfers negligible Coulomb stress

to the downdip extent of the Bannh-Dezghat thrust system (Figures 6.1 and 6.16A)while a sinistral faulting

mechanism would promote failure (Figure 6.16C). For comparison, a NW-SE trending dextral mechanism

to the Sharigh earthquake is also shown to transfer negligible Coulomb stress to the Dezghat-Bannh fault

system (Figure 6.16B). Thus, the most likely scenario for the 1931 Sharigh earthquake is rupture along a

126

NE-SW trending sinistral fault. The location of this fault near the 1997 Harnai thrust faulting earthquakes

suggest that the Sharigh earthquake lies close to the southeastern extent of the Quetta Syntaxis shear

zone.

67˚ 68˚

29˚

30˚

−150−50

−50

50

50

150

A

67˚ 68˚

150

B

67˚ 68˚

−50

−50

50

C

Figure 6.16: Amapof Coulomb stress for a receiver faultwith the samegeometry as the down-dip extensionof the Deghat-Bannh thrust fault system (gray rectangle). Contours are 50 kPa. A.) Thrust orientationfor the 1931 Sharigh earthquake. B.) Dextral orientation for the 1931 Sharigh earthquake. C.) Sinistralorientation for the 1931 Sharigh earthquake.

The northeastern end of the shear zone terminates close to the southern edge of the seismically

quiet Katawaz Block (Figure 6.1; Haq and Davis (1997)). Near the southern end of the block, immediately

north of the town of Pishin, lies the epicentral location of the 16 Nov. 1993 earthquake (Figure 6.2). In-

version of InSAR data for this earthquake shows that it ruptured a NE-SW trending, steeply dipping fault

plane in a sinistral sense similar to earthquakes elsewhere in the shear zone. The location of this earth-

quake at the southern end of the seismically quiet Katawaz block suggests that this event represents the

northwestern extent of the Quetta Syntaxis shear zone.

Using the 1993 Pishin earthquake and the 1931 Sharigh earthquake to define the NW and SE extents

of the shear zone yields a shear zone length of 100 km (Figure 6.17). Although the subsurface rupture

length of the 1931 Sharigh earthquake is unknown, using relationships from Wells and Coppersmith (1994)

and a magnitude of Mw 6.8, I estimate its subsurface rupture length as ∼ 45km. Inversion of InSAR

data for the 28 and 29 Oct. 2008 earthquakes yields subsurface rupture lengths of approximately 22 km.

Similarly, inversion of the InSAR data recording the surface deformation of the 1993 Pishin earthquake

yields a subsurface rupture length of nearly 5 km. I begin with the assumption that faults in the shear

127

zone are of uniform length, and approximate the average subsurface rupture length as 25 km. Finally,

since all of the earthquakes in the shear zone occur in the shallow crust (Table 6.2, 6.3, 6.4, 6.5), I assume

a seismogenic thickness of 15 km.

6.5.2 Shear Zone Seismic Productivity

Given an estimate of the dextral shear rate across zone the from GPS measurements, I utilize the

idealized shear zone geometry to calculate the rate of slip on each fault (Figure 6.18, Ron et al. (1984);

Sigmundsson et al. (1995)). Projecting the spatially averaged GPS velocity field (Figure 6.3) into a shear

zone-parallel geometry yields a shear zone velocity of 17.0±2.0mm/yr relative to the stable Indian Plate.

Using equations relating block rotation rates to overall shear zone velocity derived in Sigmundsson et al.

(1995), I calculate a block rotation rate of 6.8× 10−7± 0.8× 10−7rad/yr, which corresponds to a slip rate

of 9.7± 1.1mm/yr on the block-bounding faults.

Assuming the inferred slip rate and fault lengths are uniform along the shear zone, I calculate a

geometric moment of 1.8× 107 ± 0.2× 107 m3/yr, or roughly 10Mw 6.4 earthquakes per century. This

value is comparable to the observed geometric moment of 1.0× 107 m3/yr calculated from the EHB and

ISC catalogs for the past century (Table 6.1) over this same region, suggesting that the entire shear zone

has ruptured during the past century (Figure 6.19).

6.5.3 Tectonic Analogues

The southward motion of the Sulaiman Lobe relative to the stable Indian Plate, and the eastward

verging Kirthar range requires dextral shear in the Quetta Syntaxis (Figure 6.3). The absence of a promi-

nent dextral surface fault suggests that the observed dextral motion must be accommodated along con-

jugate structures. Our findings that the 1931 Sharigh earthquake, 1993 Pishin earthquake and both main-

shocks of the 2008 Ziarat earthquake sequence occurred on parallel NE/SW faults, indicates that dextral

shear is not confined to a single NW-SE trending fault. This accommodation of shear via conjugate struc-

tures is called “bookshelf faulting” and has been observed in locations as varied as the Afar, the Dead Sea,

southern California, Nicaragua and the South Iceland Seismic Zone (SISZ) (Ron et al., 1984; Garfunkel andRon,

128

?

?

Q

S

Z

P

67˚00' 67˚30' 68˚00'

30˚00'

30˚30'

15 km

Figure 6.17: Landsat 7 image from3Apr. 2001. Black lines indicatedmapped faults, andwhite lines indicatethe surface projection of faults that ruptured during the Oct.–Dec. 2008 earthquake sequence. Questionmarks are placed to indicate where the fault extent is uncertain. Faults shown with a dot-dash patternare inferred from inspection of the Landsat image as possible locations for the 1931 Sharig earthquake.Dark colors along the northern edge of the image correspond to exposed mafic and ultramafic rocks ofthe Muslimbagh ophiolite. Letters indicate the location of cities and towns: Pishin (P), Quetta (Q), Sharigh(S), and Ziarat (Z), star denotes the location of the photograph in Figure 6.14. This image is a combinationof Landsat bands 7, 4 and 2 to highlight differences in lithology.

129

w

v

L

τ

Figure 6.18: Idealized shear zone geometry, adapted from Sigmundsson et al. (1995). Blackwedges representstable boundaries to the shear zone, v is the shear velocity, L is the typical block length, w is the typicalblock width, and τ is the rotation rate.

130

1e+24

1e+25

1e+26

1e+27

1e+28

Mom

ent (d

yne−

cm

)

Chaman

Kachhi

Sharigh

Quetta

Harnai

Ziarat

NW SE

1 Fault Length2 Fault Lengths3 Fault Lengths

Cham

an F

ault

Gh

azab

and

Fau

lt Shear Zone 50 km

Figure 6.19: Moment release as a function of distance to the Quetta Syntaxis Shear Zone in fault lengths(25 km). Gray circles indicate earthquakes occurring before 1900 for which moment has been inferred.Shear zone location is indicated by the vertical dotted lines. Earthquake locations and magnitudes arefrom the EHB Centennial Catalog (Engdahl and Villasenor, 2002), historical earthquake locations (gray) arefrom Pakistan Meteorological Department and NORSAR (2007). Note that, besides the lack of locations foraftershocks to the 1931 Sharigh earthquake, the entire shear zone has ruptured in the past century.

131

1985; McKenzie and Jackson, 1986; Nicholson et al., 1986; La Femina et al., 2002; Sigmundsson, 2006). In the Afar

and the SISZ, bookshelf faulting occurs in relatively young and thin volcanic crust through the formation

of new faults, while in southern California, and Nicaragua, pre-existing structures with trends which are

conjugate to the modern axis of maximum compression are reactivated. This utilization of pre-existing

structures to accommodate overall dextral shear via conjugate sinistral faults presents a mechanically

efficient means of shearing a region (Garfunkel and Ron, 1985).

In the Quetta Syntaxis Shear Zone, the trend of the faults bounding each block varies little from

the trend of the major strike-slip faults immediately to their west (the Ghazaband and Chaman Faults,

Figure 6.17). This similarity in strike suggests that not only is the dextral shear being accommodated by

pre-existing structures, but that little clockwise rotation has occurred in this shear zone and therefore,

the onset of dextral deformation is geologically recent. That the onset of dextral deformation is recent

is in accord with the analog models of Haq and Davis (1997) that suggest the total amount of shortening

across the Sulaiman Lobe is small and that northward translation of the Katawaz Block began late in the

collision between the Indian and Eurasian Plates.

In the Dead Sea region, the age and extent of bookshelf faulting has been investigated using pa-

leomagnetically determined block rotations (Ron et al., 1984). Although no paleomagnetic measurements

exist for the western Sulaiman Lobe, a similar approach could provide an age estimate for the onset of

deformation as well as an estimate of total deformation across the Quetta Syntaxis Shear Zone.

6.6 Conclusions

The Sulaiman Lobe is a southward verging salient formed by the confined translation of a relatively

rigid Katawaz block northward by the Chaman Fault System. The overall southwardmotion of this salient

relative to a stable Indian Plate is accommodated along its eastern margin by sinistral strike-slip faulting

along structures such as the Kingri Fault. Along the western margin, overall dextral shear between the

northern Kirthar Range and the Sulaiman Lobe is accommodated by clockwise rotation of small (25 km

× 15 km) blocks through sinistral motion on NE-SW oriented block-bounding faults. The trends of these

block-bounding faults nearly parallel the strike of the Chaman Fault and likely represent reactivated faults

132

that once aided the translation of the Katawaz Block. The similarity between the strike of the block-

bounding faults and the strike of major regional faults suggests that this shear zone has experienced little

overall rotation and is therefore a young feature. Finally, the similarity between the geometric moment

release in this shear zone calculated using kinematic and seismological data suggests that the entire shear

zone has ruptured over the past century.

Chapter 7

Conclusions

7.1 Summary

I have undertaken a systematic study to determine the gross attenuation properties of the Indian

Plate using historical macroseismic data. I find, that these data are best suited to forming broad gener-

alizations about regional strong motion hazard insomuch as past seismicity can be used as a predictor of

future shaking. I also find that the low shaking attenuation observed in the Indian craton is unlike that

observed in eastern North America, an observation that is contrary to previous assumptions (Johnston,

1996; Talwani and Gangopadhyay, 2000; Ellis et al., 2001). I also find that for large cratonic earthquakes, mag-

nitudes, as estimated from calibrated attenuation parameters, are often too high. This has the effect of

biasing b-value estimations from earthquake catalogs created from macroseismic data.

I have also presented the first estimates of modern interseismic deformation rates for three lo-

cations across the Chaman Fault System, one of the least studied of the world’s major transform plate

boundaries (Figure 7.1). In addition, I have identified “bookshelf faulting” as an important moment re-

lease mechanism within the most seismically active region of the plate boundary, the Quetta Syntaxis.

Figure 7.1 shows the observed and predicted velocities for selected GPS stations with long obser-

vation durations. The predicted velocities are calculated using the pole of relative motion between the

Indian and Eurasian Plates as determined by Altamimi et al. (2007). Although the pole-of-rotation parame-

ters published by Altamimi et al. (2007) were calculated using only stations from peninsular India, the close

agreement between observed velocities along the western boundary of the Indian Plate and predicted ve-

locities (approximately 1 mm/yr residual, Figure 7.1) suggests that they provide an accurate description

134

16.8 mm/yr5.4 km 17.0 mm/yr

7.5 mm/yr2.7 km

14.7 mm/yr7.3 km

1505

18

42

1935

SIBA

KCHI

TURTNCEG

100 km

N

Figure 7.1: Summary map of interseismic deformation, as determined using space geodetic techniques,across thewestern boundary of the Indian Plate. Interseismic deformation rate and fault locking depth areindicated for three transects (black stippled rectangles) across the plate boundary and one transect acrossthe western edge of the Sulaiman Lobe (white stippled rectangle). Thin black lines represent the locationof major regional faults, thick black lines represent the approximate location of fault segments known tohave ruptured in historical times. The years of select major historical earthquakes are shown near thesegments believed to have ruptured. GPS velocities are shown in a Eurasian-Plate-fixed reference frame(Altamimi et al., 2007); the velocity of TURT is 29.96±0.42 mm/yr. Dark gray vectors represent observedGPS velocities while light gray vectors represent velocities predicted by motion about the pole of relativemotion between the Indian Plate and the Eurasian Plate. Themap is an obliqueMercator projection aboutthe pole of relative motion between the Indian and Eurasian Plates, thus points on the stable Indian Plateshow velocity vectors parallel to the lower edge of the figure. Thrust faults are shownwith black triangleson the hanging wall, all other faults are strike-slip. Fault names are indicated in Figure 4.1.

135

of themotion of the entire stable Indian Plate. Not shown in Figure 7.1 is the velocity of GPS station CHMC,

which lies within the deforming region of the plate boundary and has a small velocity relative to stable

Eurasia (approximately 6 mm/yr, see Figure 4.7).

Interseismic deformation rates across the southernmost fault in the Chaman Fault System, the

Ornach-Nal Fault, demonstrate that the obvious mud ridge west of Bela, Pakistan, is not the location of

the modern plate boundary (Figure 4.6). Instead, GPS data suggests that an unmapped fault or faults be-

neath the Hinglaj synform, west of Bela, Pakistan, are accommodating a total of 14.7 mm/yr (12.8-18.2

mm/yr 95% HPD) of interseismic deformation. This velocity indicates that the shear localized here is ap-

proximately 60% of the theoretical shear at this point on the plate boundary (Figure 4.12). The absence

of historical and modern seismicity in this region coupled with the observation of relative motion with

respect to the Indian Plate indicates that deformation must be either distributed across multiple struc-

tures, each with low deformation rates, or is manifest as surface creep. The low rates of relative motion

between the Indian Plate and sites east of the Hinglaj synform also suggests that nearly all of the diffuse

deformation associated with the plate boundary must occur west of the Hinglaj synform, on the Eurasian

Plate. This hypothesis is further supported by the observation that the westernmost GPS station on the

transect across the Ornach-Nal Fault, PANG , shows that an additional 7.4 mm/yr of fault-parallel motion

must be accommodated across the arcuate faults comprising the subaerial Makran accretionary wedge

(Figure 4.5).

Further north, at the latitude of Chaman, Pakistan, I observe the Chaman Fault to be accommo-

dating only 40% of the 19.5 mm/yr of observed relative motion with the Indian Plate. The occurrence

of the 1892 Chaman Fault earthquake, immediately south of the present day GPS transect, indicates that

the fault is seismogenic at this latitude. In addition, the shallow locking depth derived from modeling of

the GPS transect (Figure 4.8) allows for the possibility that the Chaman Fault is creeping at the surface.

This shallow locking depth, coupled with observations of seismogenic behavior (the 1892 Chaman Fault

earthquake), suggests that this segment of the Chaman Fault displays a transitional behavior between a

locked segment south of Chaman, Pakistan, and a creeping segment near the GPS transect; a behavior

similar to that observed along the Parkfield segment of the San Andreas Fault (Harris and Segall, 1987). The

136

similarity between the velocity observed at GPS station CHMC and the expected velocity in the ITRF05

reference frame (Altamimi et al., 2007) suggests that diffuse plate boundary deformationmust occur east of

the Chaman Fault. This observation is in accordance with the pattern of regional seismicity (Figure 7.2).

North of Chaman, Pakistan, the Chaman Fault veers northeastward and enters a transpressional

bend. Along the southernmost 100 km of this fault segment, the observation of interseismic deformation

rates of nearly 17 mm/yr suggests that 75% of the plate boundary deformation is accommodated by the

Chaman Fault (Figure 4.12B). Previous authors had explained the historical absence of significant seismic

moment release along this fault segment by appealing to surface creep. However, the gentle gradient

seen in the surface deformation field in Figure 4.10 precludes surface creep. This absence of fault creep

coupled with an interseismic deformation rate of nearly 17 mm/yr suggests that the Chaman Fault north

of Chaman, Pakistan, releases stored strain through seismic rupture. Additionally, the relatively low rate

of seismicity along this segment during the past 50 years (Figure 7.2) suggests that nearly 375 km of the

Chaman Fault, between Chaman, Pakistan, and Kabul, Afghanistan, represents a major seismic gap. Were

this segment to rupture in one earthquake, it could release the equivalent of an Mw > 7.6. This pos-

sibility is not accounted for in modern analyses of regional seismic hazard (Giardini et al., 1999; Pakistan

Meteorological Department and NORSAR, 2007).

The region east of the Chaman Fault near Chaman, Pakistan, accommodates the remaining 60% of

the plate boundary shear. At this latitude, the width of the deforming region of the plate boundary as

highlighted by seismicity is 150–300 km (Figure 7.2A). The exceptional seismic productivity of this region

during the 20th century relative to theChamanFault Systemproper, alongwith its complex structural style

has made this region the focus of numerous geological studies during the past half century. The complex

interplay between seismicity in theQuetta Syntaxis Shear Zone and the fold-and-thrust belts of theKirthar

Range is exemplified by the 1931 Sharigh-Mach earthquake sequence. The subsequent occurrence of the

Mw 7.7 1935 Quetta earthquake only 60 km further west adds to this complexity. I have shown, that both

seismic and macroseismic estimates of the epicentral location of the Sharigh earthquake place it within

the Quetta Syntaxis Shear Zone. Given this epicentral location, the faulting mechanism most consistent

with a stress triggering relationship to the Mach earthquake is NE-SW trending sinistral faulting.

137

66˚ 68˚ 70˚

28˚

30˚

32˚

34˚

Magnitude

6

5

4

3

A

66˚ 68˚ 70˚

1505

1842

1892

1935

1975

1978

1990

]

[

375 km

B

Figure 7.2: Seismicity and locations of historical fault rupture along the western boundary of the IndianPlate. The dashed line represents the boundary of the Katawaz Block (Chapter 4). A.) Earthquake lo-cations from 1964–2010 from the ISC catalog and moment tensors from 1976–2010 from the Global CMTProject. Filled segments of themoment tensors represent the compressional quadrants for the best-fittingdouble-couple. B.) Approximate rupture lengths for major historical strike-slip earthquakes. Rupturelengths were calculated using the relationships for strike-slip earthquakes listed in Wells and Coppersmith(1994) using published estimates of moment magnitude. The 375 km segment between the 1892 Chamanearthquake and the 1505 Kabul earthquake has no known major historical seismicity. The 1842 Jalalabadearthquake (Appendix B) was likely a thrust faulting earthquake and is shown for completeness.

138

Detailed examination of subsequent earthquakes in the Quetta Syntaxis Shear Zone, most notably

the 2008 Ziarat earthquake sequence, also indicate NE-SW trending sinistral motion. The observation

of earthquakes in this shear zone using seismic and space geodetic techniques has helped identify it as a

region of “bookshelf faulting”, similar to that found in the SanAndreas-San Jacinto fault zone aswell as the

South Iceland Seismic Zone, and Afar (Nicholson et al., 1986; Tapponnier et al., 1990; Sigmundsson et al., 1995).

Comparisons between observed seismicity throughout the shear zone over the past century combined

with calculations of a geometry-based moment budget suggests that the entire shear zone has ruptured

in the past century. The ability for this shear zone to nucleate earthquakes withMw > 6.8 combinedwith

its complete rupture during the past century suggests that this remains a region of high seismic hazard.

7.2 Future Work

The relative absence of seismotectonic analyses of the Chaman Fault System compared to otherma-

jor transform systems around the globe leavesmany fundamental questions unanswered. The harsh land-

scape and difficult security situation that sometimes exists in the region can make even basic fieldwork

difficult. Yet, the ideal natural laboratory afforded by the Chaman Fault System makes it an important

location to test hypotheses based on other transform boundaries.

The difficulty in accessing this region suggests that further research using remote sensing tech-

niques are likely to be most successful. With the soon to be ubiquitous availability of high precision, low

cost InSAR data, many locations along the Chaman Fault System will become desirable targets for study

due to the arid climate and general lack of vegetation. Although fault azimuths in southern Chaman Fault

System, in general, are incompatible with current InSAR orbits, the likelihood that faults crossing the

Hinglaj Synform are creeping at the surface make them inviting targets.

Although the southern Chaman Fault System accommodates nearly 15 mm/yr of interseismic de-

formation, modern seismic data and historical reports suggest that, at least during the past 150 years,

earthquakes are not a common occurrence along this boundary (Minchin, 1907). There remains the possi-

bility that large earthquakes, unrecorded bywritten history, have occurred in this region. The presence of

long-lived Hindu temples in the Hinglaj area could be subjected to archeological and paleoseismic studies

139

in an effort to extend the seismic record for this region to longer timescales.

Current seismotectonic hazard analysis of western Pakistan and eastern Afghanistan suggest that

most populated region along the western Indian Plate boundary to expected to experience damaging

ground accelerations in the next 50 years is Quetta, Pakistan (Giardini et al., 1999; Pakistan Meteorological

Department and NORSAR, 2007). While seismicity from the Quetta Syntaxis Shear Zone, located 30 km to

the NE, does present a threat to the city, a repeat of the Mw 7.7 1935 Quetta earthquake would be catas-

trophic. The current absence of significant seismicity along the Ghazaband Fault, the presumed source of

the 1935 Quetta earthquake, suggests that this fault could currently be accumulating strain (Figure 7.2).

However, the absence of a geodetically determined slip rate across the Ghazaband Fault, SE of Quetta pro-

hibits estimations of the strain accumulation rate, and therefore, the recurrence interval, for rupture of

this fault.

One additional benefit to increasing the spatial resolution of geodetic measurements across the

Ghazaband Fault SE of Quetta would be the resulting increase in the accuracy of convergence estimates

across the Kirthar Range. Previous geodetic studies in this region focused primarily on leveling, helping

to capture the vertical deformation caused by the 1931Mach earthquake. Triangulation across the Kirthar

Range by the Survey of India during the beginning of the 20th was performed, however, the data are not

publicly available.

Appendix A

EMS-98 Short Form

141

Table A.1: The short form of the EMS-98 intensity scale reproduced from Grunthal and Levret (2001). Fora more detailed description of the criteria used to assign intensities, refer to Grunthal and Levret (2001),specifically pages 14–20.

EMS Definition Description of typical observed effectsintensity (abstracted)

I Not felt Not felt.II Scarcely felt Felt only by very few individual people at rest in houses.III Weak Felt indoors by a few people. People at rest feel a swaying or light

trembling.IV Largely Felt indoors by many people, outdoors by very few. A few people are

observed awakened. Windows, doors and dishes rattle.V Strong Felt indoors by most, outdoors by few. Many sleeping people awake, a few

are frightened. Buildings tremble throughout.Hanging objects swing considerably. Small objects are shifted. Doors andwindows swing open or shut.

VI Slightly Many people are frightened and run outdoors. Some objects fall. Manydamaging houses suffer slight non-structural damage like hair-line cracks and fall of

small pieces of plaster.VII Damaging Most people are frightened and run outdoors. Furniture is shifted and

objects fall from shelves in large numbers. Many well built ordinaryparts buildings suffer moderate damage: small cracks in walls, fall of plaster,of chimneys fall down; older buildings may show large cracks in walls andfailure of fill-in walls.

VIII Heavily Many people find it difficult to stand. Many houses have large cracks indamaging walls. A few well built ordinary buildings show serious failure of walls,

while weak older structures may collapse.IX Destructive General panic. Many weak constructions collapse. Even well built ordinary

buildings show very heavy damage: serious failure of walls and partialstructural failure.

X Very Many ordinary well built buildings collapse.destructive

XI Devastating Most ordinary well built buildings collapse, even some with goodearthquake resistant design are destroyed.

XII Completely Almost all buildings are destroyed.devastating

Appendix B

List of Epicentral Locations for Historical Seismicity on the Indian Plate

This appendix contains one table (Table B.1) providing a summary of earthquake epicenters for

each event in the catalog for which we were able to calculate an epicenter using the methods outlined in

Chapter 3. For each earthquake, the following information is listed: column Date refers to the date of an

earthquake in the local time. A location and magnitude derived using both the minimum deviation and

minimum magnitude epicentral location methodologies described in Chapter 3 are also listed. Column

Earthquake Name provides a geographic region descriptor that links each earthquakes to the tables in the

electronic supplement toMartin and Szeliga (2010).

143Minim

umDe

viation

Minim

umMag

nitude

Earthq

uake

Date

Latit

ude

Long

itude

Depth

Mag

nitude

Latit

ude

Long

itude

Depth

Mag

nitude

Name

1762

-04-02

22.36

92.26

156.3

22.36

91.85

155.6

CHITTA

GONG

-176

218

03-09-01

28.83

78.58

157.7

29.92

78.83

157.4

BARA

HAT-

1803

1819

-06-16

23.67

70.66

158.2

22.67

70.33

157.8

ALLA

HBUN

D-18

19-A

1819

-06-17

24.07

70.08

157.2

23.07

69.99

156.2

ALLA

HBUN

D-18

19-B

1822

-04-03

24.32

90.72

157.1

22.65

88.38

156.4

BENG

AL-182

2NO

RTHIN

DIAN

1823

-02-09

7.62

84.69

157.9

6.79

81.94

156.7

OCEA

N-18

2318

33-08-26

29.18

86.52

156.6

27.6

85.35

154.7

NEPA

L-18

33-A

1833

-08-26

27.16

86.32

156.7

25.66

85.07

155.7

NEPA

L-18

33-B

1833

-08-26

27.55

86.11

157.5

27.64

85.36

156.4

NEPA

L-18

33-C

1833

-10-04

24.87

86.58

156.1

25.28

8715

5.7NE

PAL-18

33-D

1833

-10-10

25.6

85.75

156.7

26.77

83.33

155.6

NEPA

L-18

33-E

1842

-01-16

26.59

81.41

155.7

25.18

82.66

154.7

SULT

ANPU

R-18

4218

42-02-19

34.42

70.83

157.5

34.59

70.33

157.3

JALA

LABA

D-18

4218

42-03-05

30.28

80.62

157.2

30.36

78.04

154.4

GHAR

WAL

-184

218

42-11-11

24.25

88.66

157.3

22.59

88.33

156.4

BENG

AL-184

218

43-04-01

15.85

75.64

156.3

16.52

76.81

154.9

BELL

ARY-

1843

1843

-10-30

17.53

95.06

155.7

19.03

93.81

153.9

BURM

A-18

4318

45-07-23

24.85

89.9

155.9

24.01

89.23

155.4

NORT

HEA

ST-184

5-A

1845

-07-26

24.59

89.56

156.1

23.42

88.56

155.8

NORT

HEA

ST-184

5-B

1845

-08-06

26.09

90.89

157.1

23.92

89.14

156

NORT

HEA

ST-184

5-C

1846

-10-18

25.07

90.37

156.2

23.9

89.12

155.6

BENG

AL-184

618

48-04-26

NANA

NANA

NANA

NANA

SOUT

HAR

AVAL

LI-184

818

49-01-22

25.97

90.07

156.1

26.22

90.99

155.1

ASSA

M-184

918

52-03-31

28.09

79.17

157

29.92

77.83

156.2

GHAR

WAL

-185

218

56-04-06

33.51

74.27

155.6

31.6

74.43

154.8

NORT

HIND

IA-185

6-A

1856

-04-07

32.35

77.84

155.9

31.1

77.17

154.8

NORT

HIND

IA-185

6-B

1856

-04-07

32.79

74.95

156.4

31.12

77.2

155.6

NORT

HIND

IA-185

6-C

1858

-08-24

18.94

92.66

158.3

21.52

89.83

157.9

ARAK

AN-185

818

59-07-21

16.04

78.78

156.6

16.04

80.11

154.9

ONGO

LE-185

918

64-04-29

23.42

73.44

156.3

22.42

72.03

155.1

KATH

IAWAR

-186

418

65-12-15

22.3

91.05

156.8

22.71

88.63

156

CHITTA

GONG

-186

5

144Minim

umDe

viation

Minim

umMag

nitude

Earthq

uake

Date

Latit

ude

Long

itude

Depth

Mag

nitude

Latit

ude

Long

itude

Depth

Mag

nitude

Name

1865

-12-31

16.53

75.11

155.7

17.61

75.86

154.9

KARN

ATAK

A-18

6518

66-05-23

27.12

85.26

157.4

25.62

85.09

156.8

NEPA

L-18

6618

68-06-30

24.45

90.83

155.6

24.45

88.91

154.1

NORT

HEA

ST-186

818

68-07-31

22.6

85.47

155.5

23.68

86.81

154.5

JAMSH

EDPU

R-18

6818

69-01-10

24.05

93.33

158.3

25.55

91.83

157.6

CACH

AR-186

918

70-10-28

26.14

66.84

155.4

27.56

68.17

153.8

BALO

CHISTA

N-18

7018

71-01-31

23.04

73.58

156

21.13

7315

4.4SO

UTHGU

JARA

T-18

7118

78-02-05

25.95

92.3

154.8

26.12

91.72

154.5

NORT

HEA

ST-187

8-A

1878

-02-05

25.8

92.35

154.5

26.14

91.76

154.2

NORT

HEA

ST-187

8-B

1878

-03-02

34.13

74.43

157.4

33.97

72.09

156.4

HAZA

RA-187

818

78-04-29

27.65

91.48

155.2

26.57

93.98

154

NORT

HEA

ST-187

8-C

1878

-07-02

27.64

92.8

154.5

26.39

92.72

153.2

NORT

HEA

ST-187

8-D

1879

-01-03

27.88

93.81

155.5

26.8

94.23

154.4

NORT

HEA

ST-187

9-A

1879

-10-08

25.05

90.96

155.2

25.63

91.87

154.3

NORT

HEA

ST-187

9-B

1880

-06-30

25.64

93.44

155.7

26.22

91.85

155.3

NORT

HEA

ST-188

0CA

RNI

COBA

R18

81-12-31

7.17

89.6

159.3

987

.7715

8.9ISLA

ND-188

118

85-05-30

34.54

74.68

156.6

34.12

74.51

156

KASH

MIR-188

518

85-07-14

24.91

89.43

157.1

24.16

88.26

156.9

BENG

AL-188

5-A

1885

-07-14

25.68

88.91

155.1

24.43

89.66

154.4

BENG

AL-188

5-B

1897

-06-12

25.13

90.07

158.4

24.88

89.07

158.3

ASSA

M-189

7-A

1897

-06-22

24.35

88.34

156.1

25.85

89.42

155.7

ASSA

M-189

7-B

1898

-09-30

25.16

91.34

155.4

25.58

91.84

154.8

ASSA

M-189

819

00-02-08

11.03

76.58

156.4

11.37

76.83

156

COIM

BATO

RE-190

019

05-04-02

15.42

80.44

155.8

15.51

80.03

155

ONGO

LE-190

519

05-04-05

32.58

76.83

157.8

30.83

76.92

157.5

KANG

RA-190

5-A

1905

-06-13

32.05

78.36

155.6

31.13

77.2

154.2

KANG

RA-190

6-B

1905

-07-21

32.96

76.7

156.4

32.04

76.29

156

KANG

RA-190

6-C

1906

-02-28

31.73

75.36

156.9

31.14

77.2

156.2

BASH

AHR-

1906

1906

-05-20

29.68

76.56

155.9

30.35

78.06

154.9

KANG

RA-190

6-A

1906

-09-29

22.78

86.73

155.9

22.78

88.31

154.3

BENG

AL-190

619

09-10-20

28.83

68.08

05.7

29.33

67.83

05.3

KACH

HI-190

9

145Minim

umDe

viation

Minim

umMag

nitude

Earthq

uake

Date

Latit

ude

Long

itude

Depth

Mag

nitude

Latit

ude

Long

itude

Depth

Mag

nitude

Name

1912

-05-23

21.5

97.25

06.9

21.75

960

6.5BU

RMA-

1912

1915

-03-03

31.48

71.77

157.3

32.23

72.69

156

NORT

HPU

NJAB

-191

519

16-08-28

3079

.830

6.729

.4279

.420

6.3DH

ARCH

ULA-

1916

1917

-05-10

32.42

76.62

155.7

32.17

76.29

154.9

KANG

RA-191

719

18-07-08

24.73

91.23

157.4

24.23

91.65

157

SRIM

ONGA

L-19

1819

23-09-09

25.36

90.4

356.5

23.77

90.4

356.2

BANG

LADE

SH-192

319

24-01-16

18.37

73.1

155.2

19.03

72.85

153.4

MUM

BAI-1

924

1927

-06-02

22.17

80.08

357.1

25.25

8335

6.6SO

NVA

LLEY

-192

719

30-05-05

16.62

96.74

156.6

17.2

96.49

156.2

PEGU

-193

019

30-07-02

25.83

90.5

157.4

25.83

89.67

157.2

DHUB

RI-193

019

32-08-14

28.27

95.33

157

27.1

95.16

155.5

INDO

-BUR

MA-

1932

1933

-05-17

17.62

73.28

155.9

19.04

72.86

153.3

MAT

HERA

N-19

3319

34-01-11

979

.0715

5.59.4

277

.8215

2.6SIVA

KASI-193

419

34-01-15

29.19

86.09

359.3

25.52

85.34

358.6

BIHA

R-19

34-A

1934

-01-16

30.42

83.5

07.4

25.67

85.17

03.9

BIHA

R-19

34-B

1934

-01-20

27.02

85.13

06.4

25.6

85.13

05.3

BIHA

R-19

34-C

1934

-06-02

26.15

86.08

154.9

26.07

85.42

154.4

BIHA

R-19

34-D

1934

-08-29

25.86

85.83

155

26.11

85.83

154.8

BIHA

R-19

34-E

1935

-03-05

28.25

78.83

356.4

29.42

77.83

355.9

GHAR

WAL

-193

519

35-03-21

23.58

89.67

806

2489

.1780

5.9PA

BNA-

1935

1936

-02-11

2586

.7550

6.225

.9287

.5850

6BIHA

R-19

3619

36-05-27

27.18

83.86

357.2

25.85

84.61

356.8

NISH

KOTPA

HAR-

1936

1937

-10-20

31.5

78.33

356.8

30.83

76.92

355.8

HARS

IL-193

719

37-11-14

34.65

72.84

199.7

7.734

.0773

.0919

9.77.5

HIND

UKUS

H-19

3719

38-03-14

23.83

74.67

157.8

21.83

75.67

156.6

SATP

URA-

1938

1938

-04-14

20.5

94.42

158.4

24.83

89.5

156.2

BURM

A-19

3819

38-07-23

22.26

71.52

153.7

22.26

71.61

153.7

PALIAD

-193

819

38-09-10

778

.9235

6.36.8

380

.0835

5.7MAN

NAR-

1938

1941

-06-26

12.07

91.81

49.1

8.714

.5789

.9849

.18.5

ANDA

MAN

-194

119

43-10-23

26.42

93.92

07.1

25.58

91.92

06.3

ASSA

M-194

319

44-02-29

9.05

77.22

156.7

6.97

79.72

154.7

MAL

DIVE

S-19

4419

45-06-04

27.6

79.41

155.9

28.77

77.75

155

NAND

ADE

VI-194

5

146Minim

umDe

viation

Minim

umMag

nitude

Earthq

uake

Date

Latit

ude

Long

itude

Depth

Mag

nitude

Latit

ude

Long

itude

Depth

Mag

nitude

Name

1947

-07-29

27.58

93.33

06.5

27.5

94.92

04.7

SUBA

NSIRI-1

947

1950

-08-15

29.05

96.51

309

27.21

94.93

307.7

CHAY

U-19

50-A

1950

-09-13

27.07

95.03

156.8

26.16

91.78

155.4

CHAY

U-19

50-B

1951

-04-08

17.83

72.22

05.8

19.08

72.88

03.2

ARAB

IAN-

1951

1952

-10-10

30.53

69.5

05.5

30.45

69.33

05.1

BALO

CHPU

NJAB

-195

219

53-02-25

9.09

75.66

155.7

9.51

7715

4.3KE

RALA

-195

319

53-08-29

25.9

83.95

06.4

26.82

80.95

06

INDO

-NEP

AL-195

319

55-02-18

29.13

70.52

156.9

29.97

69.52

155.6

BALA

DHAK

A-19

5519

56-07-21

23.59

69.94

06.7

23.34

70.19

05.9

ANJA

R-19

5619

56-10-10

28.82

76.42

06.8

28.65

77.42

05.8

KHUR

JA-195

619

58-10-30

13.77

77.92

155.2

12.27

78.08

153.6

BANG

ALOR

E-19

5819

59-10-12

15.35

80.24

05

16.26

80.4

04.3

ONGO

LE-195

919

60-08-27

28.6

76.45

586.5

28.52

77.2

586

GURG

AON-

1960

1962

-07-13

29.98

79.11

254.9

29.56

79.61

254.5

GHAR

WAL

-196

219

64-04-15

22.6

89.4

66.9

22.1

87.9

65.2

SAUG

ORISLA

ND-196

419

65-01-12

25.39

88.1

23.6

5.426

.7288

.5223

.64.6

SIKK

IM-196

519

66-02-07

29.6

68.61

17.5

729

.8569

.5317

.56

BARK

HAN-

1966

-A19

66-02-07

32.8

70.42

156.3

30.14

71.42

154.5

BARK

HAN-

1966

-B19

66-08-01

29.8

70.13

9.87.7

30.05

68.63

9.85.5

DUKI-196

619

66-08-15

28.12

77.47

255.7

28.79

79.06

254.4

MOR

ADAB

AD-196

619

67-04-25

18.76

73.97

515

18.09

73.72

514.6

PUNE

-196

719

67-09-13

17.32

73.62

43.9

17.4

73.7

43.4

KOYN

A-19

67-A

1967

-12-10

17.22

71.85

10.7

7.717

.3973

.7710

.75.9

KOYN

A-19

67-B

1967

-12-10

17.21

72.93

105.7

17.04

74.43

104.5

KOYN

A-19

67-C

1967

-12-11

19.12

76.22

156.6

17.78

75.64

155.1

KOYN

A-19

67-D

1967

-12-24

16.76

72.63

106

17.01

74.38

104.8

KOYN

A-19

67-E

1967

-12-25

17.16

75.1

105.6

16.82

74.6

104.1

KOYN

A-19

67-F

1969

-04-13

18.31

80.92

337

17.73

80.75

336.8

BHAD

RACH

ALAM

-196

919

70-02-12

11.3

76.48

155.5

12.64

76.23

153.2

MYS

ORE-19

7019

70-07-29

28.32

92.85

79.5

7.525

.9992

.8579

.56.7

INDO

-BUR

MA-

1970

1975

-01-19

32.3

78.41

2.76.6

31.63

78.41

2.75.8

KINN

AUR-

1975

1975

-05-12

14.41

74.04

156.2

14.41

75.87

155.2

SHIM

OGA-

1975

147Minim

umDe

viation

Minim

umMag

nitude

Earthq

uake

Date

Latit

ude

Long

itude

Depth

Mag

nitude

Latit

ude

Long

itude

Depth

Mag

nitude

Name

1975

-07-08

NANA

NANA

NANA

NANA

PAGA

N-19

75BU

RMA

1977

-05-12

22.28

92.18

38.3

522

.7890

.7638

.34.8

BANG

LADE

SHBO

RDER

-197

719

80-07-29

29.7

80.93

13.8

6.829

.5480

.5213

.86.7

BAJH

ANG-

1980

1980

-11-19

NANA

NANA

NANA

NANA

SIKK

IM-198

019

84-03-20

12.55

77.43

154.8

12.63

77.77

154

BANG

ALOR

E-19

8419

86-04-26

32.23

76.32

335

32.15

76.32

334.9

DHAR

AMSA

LA-198

619

88-08-06

26.26

92.69

90.5

6.825

.1792

.6190

.56.7

INDO

-BUR

MA-

1988

1988

-08-20

27.15

86.26

577.8

26.06

86.34

577.4

UDAY

PUR-

1988

1991

-01-05

23.33

96.38

17.7

6.422

96.05

17.7

5.8MYA

NMAR

-199

119

91-01-31

36.65

70.64

142

8.133

.4872

.8914

26.9

NWFP

-199

119

91-10-20

30.57

78.39

106.5

30.74

78.56

106.2

UTTA

RKAS

HI-199

1GU

LFOF

1993

-08-25

21.15

72.19

226.1

21.56

73.02

225.3

KHAM

BAT-

1993

1993

-09-30

17.91

76.35

76.4

18.08

76.52

76.2

KILL

ARI-1

993

1993

-12-08

17.49

72.9

11.2

716

.9974

.1511

.25.2

KOYN

A-19

9319

94-02-01

16.9

74.02

105.5

17.23

73.94

105

KOYN

A-19

9419

95-12-14

19.36

76.53

105.7

18.11

76.53

102.7

MAR

ATHW

ADA-

1995

1997

-05-22

23.15

80.12

366.8

23.15

79.95

366.8

JABA

LPUR

-199

719

98-05-31

19.3

72.91

152.1

19.14

72.91

151.7

TALO

JE-199

819

99-03-29

30.98

79.51

157.3

30.4

79.01

156.4

CHAM

OLI-1

999

2000

-09-05

17.16

73.62

105.6

17.25

73.87

105

KOYN

A-20

0020

00-12-12

8.74

75.93

106.1

10.57

76.93

104.4

IDUK

KI-200

020

01-01-07

9.88

76.38

165.5

9.72

76.88

164.8

IDUK

KI-200

120

01-01-26

23.56

70.23

168.9

23.31

70.31

168.8

BHUJ

-200

1-A

2001

-01-28

24.01

70.87

126.2

23.01

72.04

125.8

BHUJ

-200

1-B

2001

-01-29

12.6

77.64

153.9

12.68

77.3

153.6

BANG

ALOR

E-20

0120

01-06-12

NANA

NANA

NANA

NANA

ORISSA

-200

1NO

RTHIN

DIAN

2001

-09-02

NANA

NANA

NANA

NANA

OCEA

N-20

0120

01-09-25

13.04

80.96

105.8

12.79

79.71

104.6

POND

ICHE

RRY-

2001

148Minim

umDe

viation

Minim

umMag

nitude

Earthq

uake

Date

Latit

ude

Long

itude

Depth

Mag

nitude

Latit

ude

Long

itude

Depth

Mag

nitude

Name

2001

-11-15

NANA

NANA

NANA

NANA

VASH

I-200

120

01-11-27

24.16

89.03

155.7

23.58

90.44

153.8

DHAK

A-20

0120

02-03-03

36.11

69.18

208.9

8.533

.8672

.8520

8.97.8

HIND

UKUS

H-20

02NO

RTH

2002

-06-20

NANA

NANA

NANA

NANA

BANG

LADE

SH-200

220

02-07-10

15.12

75.49

152.4

15.29

75.57

152.2

GADA

G-20

0220

02-11-20

35.52

74.67

134.9

35.52

74.76

134.9

ASTO

RE-200

220

03-03-10

NANA

NANA

NANA

NANA

AKOL

A-20

0320

03-03-25

26.93

90.05

50.9

626

.3488

.8850

.95.6

BHUT

AN-200

320

03-03-27

17.18

73.88

25.5

3.117

.4373

.7925

.53

KOYN

A-20

0320

03-05-27

30.44

80.01

28.9

5.230

.3678

.7628

.94.1

GHAR

WAL

-200

320

03-07-26

23.06

92.83

2.66.8

22.72

92.41

2.65.4

KOLA

BONI

A-20

0320

03-07-27

20.63

73.51

18.2

4.321

.6374

.0118

.22.3

DHAD

GAON

-200

320

03-08-05

23.15

69.58

165.6

23.06

72.58

164.2

KACH

CHH-

2003

2003

-08-10

26.55

75.65

105.5

27.96

76.31

103.5

SIKA

R-20

0320

03-12-22

NANA

NANA

NANA

NANA

CHAN

DOLI-200

320

04-01-07

12.86

71.93

156.3

14.95

74.1

151.5

GOA-

2004

2004

-02-14

34.69

73.25

35.8

6.134

.673

.3435

.86.1

ALLA

IVAL

LEY-

2004

2004

-02-20

26.71

72.21

7.95.3

25.37

72.62

7.93.7

BALO

TRA-

2004

2004

-07-06

17.65

80.5

3.93.7

17.65

80.91

3.92.6

BHAD

RACH

ALAM

-200

420

04-11-11

31.27

75.26

34.6

6.132

.1976

.3434

.64.2

KANG

RA-200

420

04-12-09

24.89

92.54

41.2

524

.892

.3741

.25

CACH

AR-200

420

04-12-22

29.32

76.46

16.1

3.729

.5776

.4616

.13.2

JIND-

2004

SUMAT

RA20

04-12-26

8.86

91.07

159.5

10.61

92.48

159.4

ANDA

MAN

-200

4OR

ISSA

BENG

AL20

05-01-01

21.64

87.18

103.4

21.64

87.35

103.1

BORD

ER-200

520

05-01-16

30.9

80.61

105.5

29.73

80.36

102.3

KUMAO

N-20

0520

05-01-30

27.56

78.53

105.1

29.31

78.44

102.3

DHAM

PUR-

2005

2005

-02-15

24.36

92.71

10.7

5.224

.8692

.810

.74.5

CACH

AR-200

520

05-02-15

31.88

73.43

59.9

5.331

.6374

.8559

.94.1

WAG

AHBO

RDER

-200

520

05-03-14

16.96

73.7

26.8

5.917

.9674

.3726

.85.7

KOYN

A-20

05-A

149Minim

umDe

viation

Minim

umMag

nitude

Earthq

uake

Date

Latit

ude

Long

itude

Depth

Mag

nitude

Latit

ude

Long

itude

Depth

Mag

nitude

Name

2005

-03-15

17.23

74.07

30.7

5.417

.5673

.9930

.75.3

KOYN

A-20

05-B

2005

-03-30

8.69

78.04

153.3

9.02

78.13

152.5

TUTICO

RIN-

2005

2005

-04-13

16.9

75.46

5.66.1

18.07

76.46

5.62.4

MAR

ATHW

ADA-

2005

2005

-06-01

NANA

NANA

NANA

NANA

ARUN

ACHA

L-20

0520

05-06-14

19.49

74.45

25.8

19.24

73.12

22.4

THAN

E-20

0520

05-07-21

21.4

91.63

11.3

5.422

.7392

.1311

.33.6

RANG

AMAT

I-200

520

05-07-24

10.58

88.82

158.8

9.16

92.07

158.6

TERE

SAISLA

ND-200

520

05-09-05

30.45

79.25

48.3

3.730

.3779

.1748

.33.6

GHAR

WAL

-200

5-A

2005

-10-08

34.69

74.06

7.98.3

32.02

75.47

7.97.6

KASH

MIR-200

5-A

2005

-10-08

NANA

NANA

NANA

NANA

KASH

MIR-200

5-B

2005

-10-15

35.05

72.92

135.9

34.05

74.08

133.7

KASH

MIR-200

5-C

2005

-10-23

34.8

74.38

12.9

5.734

.6473

.0512

.94.7

KASH

MIR-200

5-D

2005

-11-20

NANA

NANA

NANA

NANA

KOYN

A-20

05-C

2005

-11-28

20.27

86.63

105.7

21.77

87.71

102.8

SAUG

ORISLA

ND-200

520

05-12-12

36.12

70.05

210.2

8.134

.2973

.321

0.27.7

HIND

UKUS

H-20

0520

05-12-14

30.15

79.26

445.4

30.31

79.01

445.4

GHAR

WAL

-200

5-B

2005

-12-26

17.5

75.22

105.6

17.41

73.8

103.1

KOYN

A-20

05-D

2006

-01-04

16.99

7718

.64.9

18.08

76.5

18.6

2.2MAR

ATHW

ADA-

2006

2006

-02-01

29.07

81.06

9.55.3

29.9

79.81

9.53

KUMAO

N-20

06-A

2006

-02-14

28.55

88.06

306.6

27.21

88.56

304.9

SIKK

IM-200

620

06-02-24

27.38

91.7

335.6

26.21

91.7

335.2

BHUT

AN-200

620

06-03-07

24.28

69.83

156

23.62

70.75

155.1

KACH

CHH-

2006

-AKA

SHMIR

2006

-03-20

NANA

NANA

NANA

NANA

KOHI

STAN

-200

620

06-04-06

22.85

71.18

24.4

6.321

.7772

.1824

.46.1

KACH

CHH-

2006

-B20

06-04-17

17.33

73.89

11.8

4.217

.2573

.8911

.84.2

KOYN

A-20

06-A

2006

-05-07

29.96

76.81

12.8

4.628

.6377

.1412

.82.3

HARY

ANA-

2006

2006

-05-21

19.64

74.66

26.8

5.917

.4773

.6626

.82.5

KOYN

A-20

06-B

2006

-06-23

21.69

69.96

11.8

5.422

.5271

.2111

.83.2

KACH

CHH-

2006

-C20

06-08-05

24.66

89.39

104.8

23.49

89.39

103.3

BENG

AL-200

620

06-08-05

30.97

80.85

106.1

29.89

80.18

104

KUMAO

N-20

06-B

2006

-10-07

9.46

78.23

155.8

10.54

77.56

152.7

DHAR

APUR

AM-200

6

150Minim

umDe

viation

Minim

umMag

nitude

Earthq

uake

Date

Latit

ude

Long

itude

Depth

Mag

nitude

Latit

ude

Long

itude

Depth

Mag

nitude

Name

2007

-04-08

24.08

71.44

105.5

23.08

70.19

103.6

KACH

CHH-

2007

2007

-05-18

26.11

89.23

15.4

5.127

.2888

.5715

.42.2

SIKK

IM-200

7-A

2007

-05-20

27.41

88.16

42.8

3.927

.1688

.4142

.83.5

SIKK

IM-200

7-B

2007

-09-18

18.47

9332

.25.7

19.3

93.84

32.2

4.4AR

AKAN

-200

720

07-11-26

29.64

79.48

107

28.56

77.14

104.1

DELH

I-200

720

08-03-09

2370

.1710

4.723

.570

.5110

4.4KA

CHCH

H-20

0820

08-03-28

25.47

68.58

106.9

25.47

71.08

104.1

THAR

PARK

AR-200

820

08-05-29

18.14

82.92

153.8

17.64

82.67

152.7

NARS

IPAT

NAM-200

820

08-06-07

12.77

79.04

153.2

12.68

79.12

153.1

PALA

R-20

0820

08-07-30

17.89

74.92

456.3

17.47

73.67

455

KOYN

A-20

08-A

2008

-09-06

NANA

NANA

NANA

NANA

HIND

UKUS

H-20

0820

08-09-17

17.36

74.17

105.7

17.36

73.84

105.4

KOYN

A-20

08-B

2008

-10-28

30.47

67.35

155.6

30.39

67.43

155.6

CHILTA

N-20

08-A

2008

-10-29

30.77

67.63

146.4

28.18

67.88

145.7

CHILTA

N-20

08-B

2009

-02-20

34.78

75.32

106.4

34.12

73.9

105

KASH

MIR-200

9-A

2009

-03-04

NANA

NANA

NANA

NANA

HIND

UKUS

H-20

0920

09-04-09

26.22

69.83

44.3

6.526

.4771

.4244

.35.8

MOK

AL-200

920

09-05-20

34.39

76.83

156.4

33.3

75.75

154.2

KASH

MIR-200

9-B

2009

-08-11

13.26

90.01

158.7

15.85

87.92

158.2

COCO

-ISLA

ND-200

920

09-08-11

25.77

90.98

156.2

25.85

89.48

155.6

INDO

-BUR

MA-

2009

Bibliography

Altamimi, Z., X. Collilieux, J. Legrand, B. Garayt, and C. Boucher (2007), ITRF2005: A new release of the In-ternational Terrestrial Reference Frame based on time series of station positions and Earth OrientationParameters, J. Geophys. Res., 112, B09401, doi:10.1029/2007JB004949.

Ambraseys, N. (1971), Value of historical records of earthquakes, Nature, 232, 375–379.

Ambraseys, N. (2004), Three little known early earthquake in India, Curr. Sci. India, 86(4), 506–508.

Ambraseys, N., and R. Bilham (2003a), Earthquakes and associated deformation in northern Baluchistan,Bull. Seismol. Soc. Am., 93(4), 1573–1605.

Ambraseys, N., and R. Bilham (2003b), Earthquakes in Afghanistan, Seism. Res. Lett., 74(2), 107–123.

Ambraseys, N., and J. J. Douglas (2004), Magnitude calibration of north Indian earthquakes, Geophys. J. Int.,159, 165–206, doi:10.1111/j.1365-246X.2004.02323.x.

Ambraseys, N., and D. Jackson (2003), A note on early earthquakes in northern India and southern Tibet,Curr. Sci. India, 84(4), 570–582.

Apel, E., R. Burgmann, P. Bannerjee, and B. Nagarajan (2006), Geodetically constrained Indian Platemotionand implications for plate boundary deformation, Eos Trans. Am. Geophys. Union, 85(52).

Atkinson, G. M., and D. J. Wald (2007), “Did You Feel It?” intensity data: A surprisingly good measure ofearthquake ground motion, Seism. Res. Lett., 78(3), 362–368.

Babur, Z.-e.-d. M. (1912), The Babur-nama in English, Steven Austin and Sons, Hertford, translated by Bev-eredge, A. S.

Baird Smith, R. (1843a), Memoir of Indian earthquakes - part I, Journal of the Asiatic Society of Bengal, 12(136),257–293.

Baird Smith, R. (1843b), Memoir of Indian earthquakes - part II, Journal of the Asiatic Society of Bengal, 12(144),1029–1056.

Baird Smith, R. (1844), Memoir of Indian earthquakes - part III, Journal of the Asiatic Society of Bengal, 12(156),964–983.

Bakun, W. H. (1999), Erratum: Estimating earthquake location andmagnitude from seismic intensity data,Bull. Seismol. Soc. Am.

152

Bakun, W. H., and A. McGarr (2002), Differences in attenuation among the stable continental regions, Geo-phys. Res. Lett., 29(23), 2121, doi:10.1029/2002GL015457.

Bakun, W. H., and C. M. Wentworth (1997), Estimating earthquake location and magnitude from seismicintensity data, Bull. Seismol. Soc. Am., 87(6), 1502–1521.

Bakun,W. H., A. C. Johnston, andM. G. Hopper (2003), Estimating locations andmagnitudes of earthquakesin eastern North America from Modified Mercalli intensities, Bull. Seismol. Soc. Am., 93(1), 190–202.

Banks, C. J., and J.Warburton (1986), ’Passive-roof’ duplex geometry in the frontal structures of the Kirtharand Sulaiman mountain belts, Pakistan, J. Struct. Geol., 8(3/4), 229–237.

Bannert, D., A. Cheema, A. Ahmed, and U. Schaffer (1992), The structural development of theWestern FoldBelt, Pakistan, Geol. Jahrb., 80, 3–60.

Bapat, A., R. C. Kulkarni, and S. K. Guha (1983), Catalogue of earthquakes in India and neighborhood - fromhistorical period up to 1979, Indian Society of Earthquake Technology, Roorkee.

Bath, M. (1965), Lateral inhomogeneities of the upper mantle, Tectonophysics, 2(6), 483–514.

Bender, B. (1983), Maximum likelihood estimation of b values for magnitude grouped data, Bull. Seismol.Soc. Am., 73(3), 831–851.

Bender, F. K., and H. A. Raza (Eds.) (1995), Geology of Pakistan, Gebruder Borntraeger.

Berardino, P., R. Lanari, and E. Sansosti (2002), A new algorithm for surface deformationmonitoring basedon small baseline differential SAR interferograms, IEEE T. Geosci. Remote, 40(11), 2375–2383.

Bernard, M., B. Shen-Tu, W. E. Holt, and D. M. Davis (2000), Kinematics of active deformation in the Su-laiman lobe and range, Pakistan, J. Geophys. Res., 105(6), 13,253–13,279.

Bettinelli, P., J.-P. Avouac, M. Flouzat, F. Jouanne, L. Bollinger, P. Willis, and G. R. Chitrakar (2006), Platemotion of India and interseismic strain in the Nepal Himalaya from GPS and DORIS measurements, J.Geodesy, 80, 567–589, doi:10.1007/s00190-006-0030-3.

Beun, N., P. Bordet, and J.-P. Carbonnel (1979), Premieres donnees quantitatives relatives au coulissage dudecrochement de Chaman (Afghanistan du Sud-Est), C. R. Acad. Sc. Paris, 288, 931–934.

Biggs, J., T. J. Wright, L. Zhong, and B. Parsons (2007), Multi-interferogram method for measuring inter-seismic deformation: Denali fault, Alaska, Geophys. J. Int., 170, 1165–1179, doi:10.1111/j.1365-246X.2007.03415.x.

Bilham, R. (1995), Location and magnitude of the 1833 Nepal earthquake and its relation to the rupturezones of contiguous Great Himalayan eartquakes, Curr. Sci. India, 69(2), 101–128.

Bilham, R. (1998), Slip parameters for the Rann of Kachchh, India, 16 june 1819, earthquake, quantifiedfrom contemporary accounts, in Coastal Tectonics, Special Publications, vol. 146, edited by I. S. Stewart andC. Vita-Finzi, pp. 295–319, Geolocial Society, London.

Bilham, R. (2004), Earthquakes in India and the Himalaya: tectonics, geodesy and history, Annals of Geo-physics, 47, 839–858.

153

Bilham, R., and W. Szeliga (2008), Interaction between the Himalaya and the flexed Indian Plate — spatialfluctuations in seismic hazard in India in the past millenium?, in 2008 Seismic Engineering Conference com-memorating the 1908 Messina and Reggio Calabria earthquake, vol. 1020, edited by A. Santini and N. Moraci,pp. 224–231, American Institute of Physics.

Bilham, R., R. Bendick, and K. Wallace (2003), Flexure of the Indian plate and intraplate earthquakes, Proc.Indian Acad. Sci., 112(3), 315–329.

Bodin, P., L. Malagnini, and A. Akinci (2004), Ground-motion scaling in the Kachchh Basin, India, deducedfrom aftershocks of the 2001Mw 7.6 Bhuj earthquake, Bull. Seismol. Soc. Am., 94(5), 1658–1669.

Bozkurt, S. B., R. S. Stein, and S. Toda (2007), Forecasting probabilistic seismic shaking for greater Tokyofrom 400 years of intensity observations, Earthq. Spectra, 23(3), 525–546.

Burgmann, R., P. A. Rosen, and E. Fielding (2000), Synthetic aperature radar interferometry to measureEarth’s surface topography and its deformation, Ann. Rev. Earth Planet. Sci., 28, 169–209.

Burnes, A. (1835), A memoir on the eastern branch of the River Indus giving an account of the alterationsproduced by it by an earthquake in 1819, also a theory of the Runn, and some conjectures on the routeof Alexander the Great, drawn up in the years 1827–1828, Royal Asiatic Society, 3, 550–588.

Chen, W.-P., and P. Molnar (1977), Seismic moments of major earthquakes and the average rate of slip incentral Asia, J. Geophys. Res., 82(20), 2945–2969.

Clarke, P. J., D. Paradisses, P. C. England, B. Parsons, H. Billiris, G. Veis, and J.-C. Ruegg (1997), Geodeticinvestigation of the 13 May Kozani — Grevena (Greece) earthquake, Geophys. Res. Lett., 24(6), 707–710.

Day, S. M., S. H. Gonzalez, R. Anooshehpoor, and J. N. Brune (2008), Scale-model and numerical simulationsof near-fault seismic directivity, Bull. Seismol. Soc. Am., 98(3), 1186–1206.

de Montessus de Ballore, M. F. (1896), Seismic phenomena in the British Empire, Quarterly Journal of theGeological Society, 52, 651–668.

Delisle, G., U. von Rad, H. Andruleit, C. H. von Daniels, A. B. Tabrez, and A. Inam (2002), Active mud volca-noes on- and offshore eastern Makran, Pakistan, Int. J. Earth Sci., 91, 93–110, doi:10.1007/s005310100203.

Dunbar, P. K., P. A. Lockridge, and L. S. Whiteside (1992), Catalog of significant earthquakes, includingquantitative casualties and damage, Report SE-49, National Oceanic and Atmospheric Adminstration, Na-tional Environment Satellite Data and Information Service, National Geophysical Data Center.

Dziewonski, A.M., T.-A. Chou, and J. H.Woodhouse (1981), Determination of earthquake source parametersfrom waveform data for studies of global and regional seismicity, J. Geophys. Res., 86(4), 2825–2852.

Dziewonski, A. M., G. Ekstrom, and N. N. Maternovskaya (1999), Centroid-moment tensor solutions forOctober-December, 1998, Phys. Earth Planet. In., 115(1), 1–16.

Efron, B., and R. J. Tibshirani (1994), An introduction to the bootstrap, Chapman & Hall/CRC.

Ellis, M., J. Gomberg, and E. Schweig (2001), Indian earthquake may serve as analog for NewMadrid earth-quakes, Eos Trans. Am. Geophys. Union, 82(32), 345–347.

154

Engdahl, E. R., andA. Villasenor (2002), Global seismicity: 1900–1999, in International Handbook of Earthquakeand Engineering Seismology, vol. A, edited by W. H. K. Lee, H. Kanamori, P. C. Jennings, and C. Kisslinger,chap. 41, pp. 665–690, Academic Press.

Evernden, J. F., R. R. Hibbard, and J. F. Schneider (1973), Interpretation of seismic intensity data, Bull.Seismol. Soc. Am., 63(2), 399–422.

Farr, T. G., P. A. Rosen, E. Caro, R. E. Crippen, R. Duren, S. Hensley, M. Kobrick, M. Paller, E. Rodriguez,L. Roth, D. Seal, S. Schaffer, J. Shimada, J. Umland, M. Werner, M. Oskin, D. Burbank, and D. Alsdorf(2007), The shuttle radar topography mission, Rev. Geophys., 45, RG2004, doi:10.1029/2005RG000183.

Feldl, N., and R. Bilham (2006), Great Himalayan earthquakes and the Tibetian plateau,Nature, 444, 165–170,doi:10.1038/nature05199.

Frankel, A. D. (1994), Implications of felt area-magnitude relations for earthquake scaling and the averagefrequency of perceptible ground motion, Bull. Seismol. Soc. Am., 84(2), 462–465.

Freed, A. M. (2005), Earthquake triggering by static, dynamic, and postseismic stress transfer, Ann. Rev.Earth Planet. Sci., 33, 335–367, doi:10.1146/annurev.earth.33.092203.122505.

Funning, G. J., B. Parsons, and T. J. Wright (2007), Fault slip in the 1997 Manyi, Tibet earthquake fromlinear elastic modelling of InSAR displacements, Geophys. J. Int., 169, 988–1008, doi:10.1111/j.1365-246X.2006.03318.x.

Garcia, M. E., W. Szeliga, and R. Bilham (2006), Modeling vertical deformation associated with the 1931Mach earthquake, Pakistan, Eos Trans. Am. Geophys. Union, 87(52).

Garfunkel, Z., and H. Ron (1985), Block rotation and deformation by strike-slip faults 2. the properties of atype of macroscopic discontinuous deformation, J. Geophys. Res., 90(10), 8589–8602.

Giardini, D., G. Grunthal, K. M. Shedlock, and P. Zhang (1999), The GSHAP Global Seismic Hazard Map,Annali Di Geofisica, 42(6), 1225–1230.

Goldstein, R. M., and C. L. Werner (1998), Radar interferogram filtering for geophysical applications, Geo-phys. Res. Lett., 25(21), 4035–4038.

Griesbach, C. L. (1893), Notes on the earthquake in Baluchistan on the 20th December 1892, Records of theGeological Survey of India, 26(2), 57–61.

Grunthal, G., and A. Levret (Eds.) (2001), EuropeanMacroseismic Scale 1998 (EMS-98), Cahiers du Centre Europeende Geodynamique et de Seismologie, vol. 15, Joseph Beffort, Helfent-Bertrange, Luxembourg.

Gupta, S., S. S. Rai, K. S. Prakasam, D. Srinagesh, B. K. Bansal, R. K. Chadha, K. Priestley, andV. K. Gaur (2003),The nature of the crust in southern India: Implications for Precambrian crustal evolution, Geophys. Res.Lett., 30(8), 1419, doi:10.1029/2002GL016770.

Gutenberg, B., and C. F. Richter (1954), The Seismicity of the Earth and associated Phenomena, Princeton Uni-versity Press, Princeton, NJ.

Hanssen, R. F. (2001), Radar interferometry: Data interpretation and error analysis, Remote Sensing and DigitalImage Processing, vol. 2, Kluwer Academic Publishers, Dordrecht.

155

Haq, S. S. B., and D. M. Davis (1997), Oblique convergence and the lobate mountain belts of western Pak-istan, Geology, 25(1), 23–26.

Harris, R. A., and P. Segall (1987), Detection of a locked zone at depth on the parkfield, california, segmentof the san andreas fault, J. Geophys. Res., 92(8), 7945–7962, united States American Geophysical Union0885-3401 Serial.

Herring, T. A. (2002), GLOBK: Global Kalman filter VLBI and GPS analysis program, Technical report, Dep. ofEarth Atmos. and Planet. Sci. Mass. Inst. of Technol., Cambridge, Mass.

Hooper, A. (2008), Amulti-temporal InSARmethod incorporating both persistent scatterer and small base-line approaches, Geophys. Res. Lett., 35, L16302, doi:10.1029/2008GL034654.

Hough, S. E., and P. Pande (2007), Quantifying the media bias in intensity surveys: Lessons from the 2001Bhuj, India, earthquake, Bull. Seismol. Soc. Am., 97(2), 638–645, doi:10.1785/0120060072.

Hough, S. E., S. Martin, R. Bilham, and G. M. Atkinson (2002), The 26 January 2001M7.6 Bhuj, India, earth-quake obserevd and predicted ground motion, Bull. Seismol. Soc. Am., 92(6), 2061–2079.

Howell, B. F., and T. R. Schultz (1975), Attenuation of modified Mercalli intensity with distance from theepicenter, Bull. Seismol. Soc. Am., 65(3), 651–665.

Humayon, M., R. Lillie, and R. Lawrence (1991), Structural interpretation of the eastern Sulaiman foldbeltand foredeep, Pakistan, Tectonics, 10(2), 299–324.

Jackson, J., and D. McKenzie (1984), Active tectonics of the Alpine-Himalayan Belt between Turkey andPakistan, Geophy. J. R. astr. Soc., 77(1), 185–264.

Jade, S., B. C. Bhatt, Z. Yang, R. Bendick, V. K. Gaur, P. Molnar, M. B. Anand, and D. Kumar (2004), GPS mea-surments form the Ladakh Himalaya, India: Preliminary tests of plate-like or continuous deformationin Tibet, Geol. Soc. Am. Bull., 116(11/12), 1385–1391, doi:10.1130/B25357.1.

Jadoon, I. A. K., and A. Kurshid (1996), Gravity and tectonic models across the Sulaiman fold belt and theChaman fault zone in western Pakistan and eastern Afghanistan, Tectonophysics, 254, 89–109.

Jadoon, I. A. K., R. D. Lawrence, and R. Lillie (1993), Evolution of foreland structures: an example fromthe Sulaiman thrust lobe of Pakistan, southwest of the Himalayas, in Himalayan Tectonics, edited by P. J.Treloar and M. P. Searle, no. 74 in Special Publications, pp. 589–602, Geolocial Society, London, London,UK, doi:10.1144/GSL.SP.1993.074.01.39.

Jadoon, I. A. K., R. D. Lawrence, and R. J. Lillie (1994), Seismic data, geometry, evolution, and shorteningin the active Sulaiman fold-and-thrust belt of Pakistan, southwest of the Himalayas, AAPG Bull., 78(5),758–774.

Johnston, A. C. (1996), Seismic moment assessment of earthquakes in stable continental regions — II. his-torical seismicity, Geophys. J. Int., 125(3), 639–678.

Jones, A. (Ed.) (1961), Reconnaissance Geology of part of West Pakistan, a Columbo Plan cooperative project, Hunt-ing Survey Report, Government of Pakistan, Oshawa.

Joyner, W. B., and D. M. Boore (1993), Methods for regression analysis of strong-motion data, Bull. Seismol.Soc. Am., 83(2), 469–487.

156

Khan, K. (1998), Determination of focal mechanisms and hypocenters for the mainshock and aftershocksof the 1997 Quetta earthquake in Pakistan, Individual Studies by Participants to the International Institute ofSeismology and Earthquake Engineering, 34, 131–139.

Khattri, K. N. (1987), Great earthquakes, seismicity gaps and potential for earthquake disaster along theHimalayan plate boundary, Tectonophysics, 138, 79–92.

Khattri, K. N., and A. K. Tyagi (1983), Seismicity patterns in the Himalayan plate boundary and identifica-tion of the areas of high seismic potential, Tectonophysics, 96, 281–297.

King, R. W., and Y. Bock (2002), Documentation for the GAMIT GPS analysis software, Technical report, Dep.of Earth Atmos. and Planet. Sci. Mass. Inst. of Technol., Cambridge, Mass.

Kozuch, M. J. (1995), Earthquake hazard analysis of Venezuela using site specific attenuation, Ph.D. thesis,University of Colorado.

Kumar, S., S. G. Wesnousky, T. K. Rockwell, R. W. Briggs, V. C. Thakur, and R. Jayangondaperumal (2006),Paleoseismic evidence of great surface rupture earthquakes along the Indian Himalaya, J. Geophys. Res.,111, B03304.

La Femina, P. C., T. H. Dixon, andW. Strauch (2002), Bookshelf faulting inNicaragua, Geology, 30(8), 751–754.

Landor, A. H. S. (1902), Across coveted lands or a Journey from Flushing (Holland) to Calcutta Overland, Macmillanand Co., London.

Lave, J., D. Yule, S. Sopkota, K. Basant, C. Madden, M. Attal, and R. Pandey (2005), Evidence for a GreatMedieval earthquake (∼1100 A.D.) in the central Himalayas, Nepal, Science, 307, 1302–1305.

Lawrence, R. D., and R. S. Yeats (1979), Geological reconnaissance of the Chaman Fault in Pakistan, inGeodynamics of Pakistan, edited by A. Farah and K. A. DeJong, Geological Survey of Pakistan, Quetta.

Lawrence, R. D., S. H. Khan, and T. Nakata (1992), Chaman Fault, Pakistan-Afghanistan, Annales Tectonicae,6, 196–223.

Lenox-Conyngham, G. P. (1916), Leveling of precision in India, Addendum, Heights of bench marks, vol. Sheet 39,Survey of India, Dehra Dun, India.

Lienkaemper, J. J., B. Baker, and F. S. McFarland (2006), Surface slip associated with the 2004 Parkfield,California, earthquake measured with alinement arrays, Bull. Seismol. Soc. Am., 96(4), 239–249, doi:10.1785/0120050806.

Lohman, R. B., and M. Simons (2005), Some thoughts on the use of InSAR data to constrain models ofsurface deformation: Noise structure and data downsampling, Geochem. Geophys. Geosyst., 6(1), Q01007,doi:10.1029/2004GC000841.

MacMurdo, J. (1823), Papers relating to the earthquake which occurred in India in 1819, Literary Society ofBombay Transactions, 3, 90–116.

Malik, J. N., T. Nakata, H. Sato, T. Imaizumi, T. Yoshioka, G. Philip, A. K. Mahajan, and R. V. Karanth (2001),January 26, 2001, the Republic Day (Bhuj) earthquake of Kachchh and active faults, Gujarat, westernIndia, Journal of Active Fault Research, Japan, 20, 112–126.

157

Mallet, R., and J. W. Mallet (1858), The earthquake catalogue of the British Association with the discussion, curves,and maps, etc., Transactions of the British Association for the advancement of science, 1852 to 1858,Taylor and Francis, London.

Manighetti, I., M. Campillo, C. Sammis, P. M. Mai, and G. King (2005), Evidence for self-similar, triangularslip distributions on earthquakes: Implications for earthquake and fault mechanics, J. Geophys. Res., 110,B05302, doi:10.1029/2004JB003174.

Martin, S., and W. Szeliga (2010), A catalog of felt intensity data for 589 earthquakes in India from1636–2009, Bull. Seismol. Soc. Am., 100(2), 562–569, doi:10.1785/0120080328.

Massonnet, D., and K. Feigl (1995), Discrimination of geophysical phenomena in satellite radar interfero-grams, Geophys. Res. Lett., 22(12), 1537–1540.

McCaffrey, R., G. A. Abers, and P. C. Zwick (2000), Inversion of teleseismic body waves, in Digital Seismo-gram Analysis and waveform inversion, IASPEI Software Library, vol. 3, edited by W. H. K. Lee, InternationalAssociation of Seismology and Physics of the Earth’s Interior.

McKenzie, D., and J. Jackson (1986), A blockmodel of distributed deformation by faulting, J. Geol. Soc. London,143, 349–353.

Medvedev, S. V., W. Sponheuer, and V. Karnik (1965), Seismic intensity scale versionMSK 1964, Moscow, Akad.Nauk SSSR Geofiz. Kom., Moscow.

Miao, Q., and C. A. Langston (2008), Comparative study of distance attenuation in Central United Statesand Western India, Seism. Res. Lett., 79(3), 446–456.

Middlemiss, C. S. (1910), The Kangra earthquake of 4th April 1905, Memoirs of the Geological Survey of India,38.

Minchin, C. F. (Ed.) (1907), The Gazatteer of Las Bela, District Gazatteer series of Baluchistan, vol. 8, The TimesPress, Bombay.

Mitra, S., K. F. Priestley, V. K. Gaur, and S. S. Rai (2006), Frequency-dependentLg attenuation in the Indianplatform, Bull. Seismol. Soc. Am., 96(6), 2449–2456, doi:10.1785/0120050152.

Molnar, P., and J. M. Stock (2009), Slowing of India’s convergence with Eurasia since 20 Ma and its impli-cations for Tibetan mantle dynamics, Tectonics, 28, TC3001, doi:10.1029/2008TC002271.

Molnar, P., and P. Tapponnier (1975), Cenozoic tectonics of Asia: Effects of a continental collision, Science,189(4201), 419–426.

Monsalve, G., A. F. Sheehan, C. Rowe, and S. Rajaure (2008), Seismic structure of the crust and uppermantlebeneath the Himalayas: Evidence for eclogitization of lower crustal rocks in the Indian Plate, J. Geophys.Res., 113, B08315, doi:10.1029/2007JB005424.

Mosegaard, K., and A. Tarantola (1995), Monte Carlo sampling of solutions to inverse problems, J. Geophys.Res., 100(7), 12,431–12,447.

Nicholson, C., L. Seeber, and P. Williams (1986), Seismic evidence for conjugate slip and block rotationwithin the San Andreas fault system, southern California, Tectonics, 5(4), 629–648.

158

Okada, Y. (1992), Internal deformation due to shear and tensile faults in a half-space, Bull. Seismol. Soc. Am.,82(2), 1018–1040.

Oldham, R. D. (1899), Report of the Great Earthquake of 12th June 1897, Memoirs of the Geological Survey ofIndia, 29.

Oldham, R. D. (1926), The Cutch (Kachh) earthquake of 16th June 1819 with a revision of the great earth-quake of 12th June 1897, Memoirs of the Geological Survey of India, 46, 71–177.

Oldham, T. (1883), A catalogue of Indian earthquakes from the earliest time to the end of A.D. 1869,Memoirsof the Geological Survey of India, 29, 163–215.

Ortiz, M., and R. Bilham (2003), Source area and rupture parameters of the 31 December 1881 Mw = 7.9Car Nicobar earthquake estimated from tsunamis recorded in the Bay of Bengal, J. Geophys. Res., 108(4),2215, doi:10.1029/2002JB001941.

Pakistan Meteorological Department, and NORSAR (2007), Seismic hazard analysis and zonation for Pak-istan, Azad Jammu and Kashmir, Tech. rep., Pakistan Meteorological Department.

Pande, P., and J. R. Kayal (Eds.) (2003), Kutch (Bhuj) earthquake 26 January 2001, no. 76 in Special Publications,Geological Survey of India, Kolkata.

Parsons, B., T. J. Wright, P. Rowe, J. Andrews, J. Jackson, R. Walker, M. Khatib, M. Talebian, E. Bergman,and E. R. Engdahl (2006), The 1994 Sefidabeh (eastern Iran) earthquake revisited: new evidence fromsatellite radar interferometry and carbonate dating about the growth of an active fold above a blindthrust fault, Geophys. J. Int., 164, 202–217, doi:10.1111/j.1365-246X.2005.02655.x.

Pasyanos, M. E., E. M. Matzel, W. R. Walter, and A. J. Rodgers (2009), Broad-band Lg attenuation modellingin the Middle East, Geophys. J. Int., 177(3), 1166–1176, doi:10.1111/j.1365-246X.2009.04128.x.

Powell, M. J. D. (1964), An efficient method for finding the minimum of a function of several variableswithout calculating derivatives, Comput. J., 7(2), 152–162.

Quittmeyer, R., A. Farah, and K. H. Jacob (1979), The seismicity of Pakistan and its relation to surface faults,inGeodynamics of Pakistan, edited byA. Farah andK. A. DeJong, pp. 271–284, Geological Survey of Pakistan.

Quittmeyer, R., A. L. Kafka, and J. G. Armbruster (1984), Focal mechanisms and depths of earthquakes incentral Pakistan: a tectonic interpretation, J. Geophys. Res., 89, doi:10.1029/JB089iB04p02459.

Raghu Kanth, S. T. G., and R. N. Iyengar (2006), Seismic hazard estimation for Mumbai city, Curr. Sci. India,91(11), 1486–1494.

Rajendran, C. P., and K. Rajendran (2001), Characteristics of deformation and past seismicity associatedwith the 1819 Kutch earthquake, northwestern India, Bull. Seismol. Soc. Am., 91(3), 407–426.

Rajendran, C. P., and K. Rajendran (2005), The status of central seismic gap: a perspective based on thespatial and temporal aspects of the large Himalayan earthquakes, Tectonophysics, 395, 19–39.

Rajendran, C. P., K. Rajendran,M. Thakkar, andB. Goyal (2008), Assessing theprevious activity at the sourcezone of the 2001 Bhuj earthquake based on the near-source and distant paleoseismological indicators,J. Geophys. Res., 113, B05311, doi:10.1029/2006JB004845.

159

Ramachandran, K., and L. S. Srivastava (1991), New catalogue of felt Indian earthquakes during 1901–1971,Mausam, 42(2), 171–182.

Ramanathan, K. R., and S.M.Mukherji (1938), A seismological study of the Baluchistan (Quetta) earthquakeof May 31, 1935, Records of the Geological Survey of India, 73(4), 483–513.

Ron, H., R. Freund, and Z. Garfunkel (1984), Block rotation by strike-slip faulting: structural and paleo-magnetic evidence, J. Geophys. Res., 89(7), 6256–6270.

Rosen, P. A., S. Hensley, G. Peltzer, and M. Simons (2004), Updated Repeat Orbit Interferometry packagereleased, Eos Trans. Am. Geophys. Union, 85(5), 35.

Rowlands, D. (1978), The structure and seismicity of a portion of the southern Sulaiman Range, Pakistan,Tectonophysics, 51, 41–56.

Savage, J. C., and R. O. Burford (1973), Geodetic determination of relative platemotion in central California,J. Geophys. Res., 78(5), 832–845.

Schelling, D. (1999a), Structural geology of the Bolan block, western Pakistan, EGI Technical Report 5-20885-99, Energy and Geoscience Institute, University of Utah, Salt Lake City, Utah.

Schelling, D. (1999b), Geological map of the Bolan Block and Quetta Syntaxis regions, EGI Technical Report5-20885-99, Energy and Geoscience Institute, University of Utah, Salt Lake City, Utah.

Schmidt, D. A., and R. Burgmann (2006), InSAR constraints on the source parameters of the 2001 Bhujearthquake, Geophys. Res. Lett., 33, L02315, doi:10.1029/2005GL025109.

Seeber, L., and J. G. Armbruster (1981), Great detachment earthquakes along the Himalayan arc and long-term forecasting., in Earthquake Prediction — An International Review,Maurice Ewing Series, vol. 4, edited byD. W. Simpson and P. G. Richards, pp. 259–277, American Geophysical Union.

Segall, P., and J. L. Davis (1997), GPS applications for geodynamics and earthquake studies, Ann. Rev. EarthPlanet. Sci., 25, 301–336.

Sigmundsson, F. (2006), Iceland Geodynamics Crustal Deformation and Divergent Plate Tectonics, chap. 7, p. 232,Springer Praxis Books, Springer Berlin Heidelberg, doi:10.1007/3-540-37666-6.

Sigmundsson, F., P. Einarsson, R. Bilham, and E. Sturkell (1995), Rift-transform kinematics in south Ice-land: deformation from global positioning system measurements, 1986 to 1992, J. Geophys. Res., 100(4),6235–6248.

Singh, D. D., and H. K. Gupta (1980), Source dynamics of two great earthquakes of the Indian subcontinent:the Bihar-Nepal earthquake of January 15, 1934 and the Quetta earthquake of May 30, 1935, Bull. Seismol.Soc. Am., 70(3), 757–773.

Singh, S. K., R. S. Dattatrayam, and H. K. Gupta (1999), A spectral analysis of the 21 May 1997, Jabalpur,India, earthquake (Mw = 5.8) and estimation of ground motion from future earthquakes in the Indianshield region, Bull. Seismol. Soc. Am., 89(6), 1620–1630.

Snead, R. E. (1964), Active mud volcanoes of Baluchistan, west Pakistan, Geogr. Rev., 54(4), 546–560.

Srivastava, H. N., and K. Ramachandran (1985), New catalogue of earthquakes for peninsular India during1839–1900, Mausam, 36(3), 351–358.

160

Stein, R. S., G. C. P. King, and J. Lin (1994), Stress triggering of the 1994 Mw = 6.7 Northridge, California,earthquake by its predecessors, Science, 265(5177), 1432–1435.

Talwani, P., and A. Gangopadhyay (2000), Schematic model for intraplate earthquakes, Eos Trans. Am. Geo-phys. Union, 81(48), S21C–10.

Tandon, A. N., and H. M. Chaudhury (1968), Koyna earthquake of December, 1967, Sci. Rep 59, India Meteo-rol. Dept.

Tapponnier, P., R. Armijo, I. Manighetti, and V. Courtillot (1990), Bookshelf faulting and horizontal blockrotations between overlapping rifts in southern Afar, Geophys. Res. Lett., 17(1), 1–4.

Treloar, P. J., and M. P. Coward (1991), Indian plate motion and shape: constraints on the geometry of theHimalayan orogen, Tectonophysics, 191, 189–198.

Wald, D. J., and T. I. Allen (2007), Topographic slope as a proxy for seismic site conditions and amplification,Bull. Seismol. Soc. Am., 97(5), 1379–1395, doi:10.1785/0120060267%20.

Wald, D. J., V. Quitoriano, L. Dengler, and J. Dewey (1999a), Utilization of the Internet for rapid communityintensity maps, Seism. Res. Lett., 70(6), 680–697.

Wald, D. J., V. Quitoriano, T. H. Heaton, and H. Kanamori (1999b), Relationships between peak groundacceleration, peak ground velocity, and Modified Mercalli Intensity in California, Earthq. Spectra, 15(3),557–564.

Waldhauser, F., and W. L. Ellsworth (2000), A double-difference earthquake location algorithm: Methodand application to the northern Hayward Fault, California, Bull. Seismol. Soc. Am., 90(6), 1353–1368.

Wellman, H. W. (1966), Active wrench faults of Iran, Afghanistan and Pakistan, Geol. Rundsch., 55(3),716–735, doi:10.1007/BF02029650.

Wells, D. L., and K. J. Coppersmith (1994), New empirical relationships among magnitude, rupture length,rupture width, rupture area, and surface displacement, Bull. Seismol. Soc. Am., 84(4), 974–1002.

Wessel, P., and W. H. F. Smith (1998), New improved version of Generic Mapping Tools released, Eos Trans.Am. Geophys. Union, 79, 579.

West, W. D. (1934), The Baluchistan earthquakes of August 25th and 27th, 1931, Memoirs of the GeologicalSurvey of India, 67, 82.

Wilson, C. A. (1938), Leveling operations, Technical Report 1937, Survey of India, Dehra Dun, India.

Wright, T. J., B. Parsons, and E. Fielding (2001), Measurement of interseismic strain accumulation acrossthe North Anatolian Fault by satellite radar interferomety, Geophys. Res. Lett., 28(10), 2117–2120.

Wyss, M. (1979), Estimating maximum expectable magnitude of earthquakes from fault dimensions, Geol-ogy, 7(7), 336–340.

Yeats, R. S., R. D. Lawrence, S. Jamil-ud din, and S. H. Khan (1979), Surface effects of the 16 March 1978earthquake, Pakistan-Afghanistan border, in Geodynamics of Pakistan, edited byA. Farah andK. A. DeJong,Geological Survey of Pakistan, Quetta.

Zaigham, N. A. (1991), Bela ophiolites and associated mineralization in southern part of axial-belt of Pak-istan, Ph.D. thesis, University of Karachi.