Вы находитесь на странице: 1из 67

AN ABSTRACT OF THE THESIS OF

Emily M. Oatney for the degree of Master of Science in Geology presented on October 9,
1998. Title: Geology and Paleoseismology of the Trans-Yamuna Active Fault System,
Himalayan Foothills of Northwest India.

Redacted for privacy


Abstract approved:
Robert S. Yeats

Satellite image interpretation, geologic mapping, and paleoseismic trenching are

used to investigate the Trans-Yamuna active fault system in the northwestern Doon Valley
of the Indian Himalayan foothills. This east-west fault system is subparallel to and crosses
the Main Boundary thrust near the structural transition from the Nahan salient to the

Dehra Dun reentrant. The Trans-Yamuna active fault system may terminate to the east at
a lateral ramp of the Main Boundary thrust. A south-side-up, relatively linear fault trace
with variable fault dips suggests that the fault system is high-angle reverse with a
component of strike-slip. It is subdivided into the Sirmurital, Dhamaun, and Bharli faults,

which probably connect at depth. The Dhamaun fault is exposed where it cuts the late
Holocene upper Bhatrog terrace deposit of the Giri River. A paleoseismic investigation of
the Sirmurital fault at another Giri River terrace did not expose the fault, but it suggests

that late Holocene terrace deposits there may be folded into a syncline parallel to fault
strike. The fold axis of the syncline continues into bedrock to the west. Earthquakes in

1905, 1803, or perhaps earlier may have been the source of folding of the fine-grained
sediments within this terrace deposit. The Trans-Yamuna active fault system is a
secondary hangingwall fault that may accommodate some strain release above the

decollement during large-magnitude earthquakes. Strike-slip motion may be related to the


lateral translation of the Karakoram fault block and east-west extension of the southern
Tibet block as a result of oblique convergence between the Indian and Eurasian plates in
the northwest Himalaya.

Copyright by Emily M. Oatney

October 9, 1998
All Rights Reserved

Geology and Paleoseismology of the Trans-Yamuna Active Fault System,


Himalayan Foothills of Northwest India

by

Emily M. Oatney

A THESIS
submitted to

Oregon State University

in partial fulfillment of

the requirements for the

degree of

Master of Science

Presented October 9, 1998


Commencement June 1999

Master of Science thesis of Emily M. Oatney presented on October 9, 1998

APPROVED:

Redacted for privacy


Major Professor, representing/Geology

Redacted for privacy


Chair of Department of Geosciences

Redacted for privacy


Dean of Graduate School

I understand that my thesis will become part of the permanent collection of Oregon State
University libraries. My signature below authorizes release of my thesis to any reader
upon request.

Redacted for privacy


Emily M. Oatney, Author

ACKNOWLEDGMENTS

This study was funded by National Science Foundation Grants EAR- 93-03577

and EAR-97-25208 as part of a collaboration between Oregon State University (OSU)


and the Wadia Institute of Himalayan Geology (WIHG) in Dehra Dun, India. The Oil and
Natural Gas Corporation of India (ONGC) was also involved in this study under a
separate memorandum of understanding with OSU.

I would like to thank Bob Yeats for making this project possible, for allowing me
to live and work in India, and for guidance and support through all stages of the project.
I thank WHIG for project cooperation and help with fieldwork logistics, especially

Dr. V.C. Thakur, director, and Dr. N.S. Virdi, who served as my thesis advisor while I
worked in India. Special thanks are given to Sh. Shayam Singh for excellent driving in

challenging terrain. Dr. N.S. Virdi, Sh. A. Pandey, Dr. M.P. Sah, and Dr. G. Philip helped
clean and log the Sirmurital trench. I would also like to thank the contract trench
excavators of Sirmurital village for three months of digging with limited resources.

Tom Fumal at the United States Geological Survey and Ray Weldon at the

University of Oregon advised me about how to log and photograph the trench using their
`Trench-o-matic' camera frame design.

I thank ONGC for supporting the goals of this study, and the Remote Sensing

Group of ONGC for processing the digital satellite data used for interpretation. K.K.S.
Pangthey provided field insights on bedrock structure. I am especially grateful to Firoze
and Sucheta Dotiwala, and Ram "Navi" Kumar, who made my stay in Dehra Dun a
lifetime experience never to be forgotten. In addition to supporting my affiliation and
work with both ONGC and WIHG, they showed me the beauty of Indian culture and
helped me to fit into the local community.

Lastly, I would like to thank my parents, my brother Mark, and many fellow

students and friends for encouragement and support throughout my stay at Oregon State
University.

TABLE OF CONTENTS
Page
INTRODUCTION

METHODS

GEOLOGIC SETTING

Tectonic Setting of the Himalaya


Regional Earthquake History
Stratigraphic Setting of the Himalayan Foothills
Doon Valley Structure

TRANS-YAMUNA GEOLOGY

7
8
12
15

18
18

Local Bedrock Stratigraphy and Structure


Quaternary Deposits and Geomorphology

22

TRANS-YAMUNA ACTIVE FAULT SYSTEM

25

Sirmurital Fault
Dhamaun Fault
Bharli Fault
Other Active Faults

26
26
29
32

PALEOSEISMIC INVESTIGATION AT SIRMURITAL

36

Site Description
Trench Stratigraphy
Radiocarbon Dating
Quaternary Synclinal Folding at Sirmurital

36
37
39
40

DISCUSSION
Style of the Trans-Yamuna Active Fault System
Terrace Ages and Folding as Indicators of Fault Timing
Influence of Oblique Convergence Between India and Eurasia
Large Magnitude Earthquake Recurrence and Seismic Hazard

44
44
45
46
48

CONCLUSIONS

50

REFERENCES CITED

52

LIST OF FIGURES
Page

Figure
1.

Tectonic map of the western Himalaya

2.

Crustal-scale generalized Himalayan cross section showing ramp and flat


structure and concentrated seismicity

Rupture zones of the great 1897, 1905, 1934, and 1950 Himalayan
earthquakes (shaded areas)

10

4.

Meisoseismal regions of the 1905 Kangra and 1803 Mathura earthquakes

11

5.

Regional stratigraphy of the Himalayan foothills of India

13

6.

Generalized geologic map of the Doon Valley region, northwest India

16

7.

Landsat Thematic Mapper satellite image of the Trans-Yamuna active fault


system, northwest Doon Valley, India

19

8.

Topographic map of the Trans-Yamuna area with faults and fluvial terraces

24

9.

Photograph of Sirmurital terraces (far side) and Bhatrog terraces (near side)
of the Giri River

27

Photograph of the Dhamaun fault (yellow arrows) exposed along the Giri
River

28

Photograph of entrained Giri River terrace pebbles (p) in sheared Dagshai


Formation along the Dhamaun fault

30

12.

Photograph of a sag pond on the north side of Bharli fault

31

13.

Photographs of a Bharli fault exposure

33

14.

Photograph and sketch of an active fault at Kalawar Village with 40 m of


apparent right-lateral separation

35

Photograph of Sirmurital trench, showing stationary camera frame


and shoring device

38

16.

Photographs of south-dipping strata in the lower Sirmurital terrace

41

17.

Convergence diagram of India and north Tibet, showing arc-parallel


convergence south of the Karakoram fault

47

3.

10.

11.

15.

LIST OF PLATES
Plate
1.

Geologic Map of the Trans-Yamuna Area, Northwestern Doon Valley, India

2.

Sirmurital Trench Log, Northwestern Doon Valley, India

GEOLOGY AND PALEOSEISMOLOGY OF THE TRANS-YAMUNA ACTIVE


FAULT SYSTEM, HIMALAYAN FOOTHILLS OF NORTHWEST INDIA

INTRODUCTION

The Himalayan mountain range is the largest active continent-continent collision

zone in the world. It is located where the Indian plate slips beneath the Eurasian plate,
resulting in a doubled crustal thickness (Molnar, 1987; Hirn et al, 1984) and in thrust

stacks of Indian crust and cover rocks that reach altitudes greater than 8 km (Gansser,
1964; Seeber et al., 1981). The Himalaya is subject to great earthquakes, which nucleate
at depth along the shallow north-dipping decollement and propagate southward beneath
the heavily-populated Himalayan foothills to the deformation front (Yeats and Lillie, 1991)

(Figs. 1, 2).
Four major earthquakes with Mw 8 or greater, in 1897, 1905, 1934, and 1950,
ruptured approximately 50 % of the Himalayan front length over the past century (Seeber

and Armbruster, 1981; Molnar, 1990). Geology, geodesy, seismicity, and historical
records together create a basis for hazard assessment in seismic gaps along the Himalayan

front. Long-term geologic and short-term geodetic convergence rates indicate that the
Himalayan earthquake recurrence intervals are probably less than 500 years in Nepal

(Bilham, 1995; Bilham et al., 1998). Careful documentation of Holocene deformation


supplements earthquake historical records, especially where written accounts are sparse.

In northwestern India, the only great earthquake recorded in written accounts during the
last 1000 years is the 1905 Kangra earthquake. It is unclear if there was a hiatus of large
earthquakes, or if they were simply not recorded in temple or gazette records (Kapur,
1934).

An important factor in seismic hazard evaluation of the Himalaya is whether all of

the accumulating strain above the locked decollement is released at the deformation front

or distributed over a broader area. In Nepal, geodetic and geologic convergence rates are
similar, which supports the idea that strain release there is focused at the front (Bilham et

36

INDIA
Jhelum
Syntax's

32
HFT

Lahore!

\ TIBET

Nahan

Salient

PAKISTAN
30

Fig. 6

NEPAL

INDIA
Dehra Dun
Reentrant

Delhi
28
N

500 km

72 E

74

76

78

80

82

84

Figure 1. Tectonic map of the western Indian Himalaya. Dominant regional faults
include the Indus-Tsangpo suture zone (ITSZ), Karakoram fault, Main Central
thrust (MCT), Main Boundary thrust (MBT), Main Mantle thrust (MMT), and
Salt Range thrust (SRT). The Nahan salient and Dehra Dun reentrant are shown
south of the MBT. The Himalayan Front fault (HFT), at the southern boundary
of the Sub-Himalaya (shown as stippled pattern), is not continuous at the surface.
National boundaries are shown with dashed lines. The highlighted box is detailed in
Figure 6. Modified from Powers et al. (1998).

focus of this study

Lesser Himalaya
0

Indo-Gangetic Plains HPT

4487-

Sub-Himalaya

doe-

High Himalaya
-.1MM" AhIPP''1111Wilim.

INDIAN CRUST

50

50

100

150

200

250

Undeformed foreland strata

Himalayan crystalline basement

Active foreland fold-and-thrust belt

Earthquake hypocenters with


preferred fault plane solutions

Deformed-passive margin strata

300

Figure 2. Crustal-scale generalized Himalayan cross section showing ramp and


flat structure and concentrated seismicity. The main Himalayan thrust faults,
from south to north, are the Himalayan Front thrust (HFT), the Main Boundary
thrust (MBT), the Main Central thrust (MCT), and the Main Himalayan fault
(MHT) at depth. From Powers et al. (1998) with structural and seismic
information from Ni and Barazangi (1984), Schelling and Arita (1991), Zhao
et al. (1993), and Jackson and Bilham (1994).

al., 1998). This may not be the case in the western Himalaya, where geodetic and
geologic convergence rates are not as well constrained.
The Doon Valley of northwestern India was chosen as the region within the
Himalayan foothills to search for active faults because of the presence of prominent
lineaments identified on a published Landsat Thematic Mapper (TM) satellite image

(Philip, 1996). The south-side-up offset of these fault lineaments is similar to that of
active faults in Nepal along the Main Boundary thrust (Nakata, 1982; 1986; 1989).
The objective of this study is to identify and characterize active faults and related

Holocene deformation within the Trans-Yamuna area of Himachal Pradesh state, India,
using satellite image interpretation, geologic mapping, and paleoseismic trenching. The
trench described in this study is the first paleoseismic trench excavation in the Himalaya.
Mapping and trenching results presented in this study indicate that lineaments identified on

satellite imagery are part of an active fault system in the Trans-Yamuna area that may
contribute to western Himalayan convergence rates at and north of the Himalayan front.

METHODS

Digital Landsat TM satellite imagery at 30 m resolution was used to initially

identify and map fault lineaments and Quaternary deposits in the Trans-Yamuna area. The
images, provided by the Indian Oil and Natural Gas Corporation (ONGC), were processed

in the company's remote sensing laboratory. Indian Remote Sensing (IRS-C)


panchromatic (PAN) digital satellite imagery at 5.8 m resolution was available toward the
end of the study, and was used to confirm fault placement and morphology within the
study area during the final compilation of field data.

The topographic maps for the Trans-Yamuna area, used as a base for geologic
mapping, are available from the Survey of India in Dehra Dun at a scale of 1:50,000 and a

contour interval of 40 m. The geologic map produced in this study covers 110 km2 and is
presented at a scale of 1:35,000. An emphasis was placed on Quaternary tectonic features
associated with the Trans-Yamuna active fault system. Most field work was completed
during the post-monsoon autumn seasons of 1996 and 1997.
The paleoseismic trench investigation described in this study was carried out in the

autumn of 1996 and early winter of 1997. Bulldozers were available for excavation, but
they were not appropriate to use at the trench site because of access and agricultural land
use issues. Five men from Sirmurital village were hired over a period of three months to
hand dig the trench using pick axes and shovels. Wadia Institute of Himalayan Geology

(WIHG) organized the labor contracts. The Sirmurital trench was oriented perpendicular
to the scarp and was about 10 m long, 4 m wide, and 4 m deep, with vertical walls. A
fence was put up around the trench for safety.
The north wall of the trench collapsed after a storm, before shoring devices were
placed. To brace the remaining wall, planks were placed vertically against the wall with

eucalyptus logs wedged in between it and the other side of the trench. Hydraulic shores
were not available. In retrospect, rather than using unstable plank and log shoring
devices, the trench could have been dug in a benched or 45 laid back style. This would
have increased the time of digging, but safety issues would have been minimized.

Standard cleaning and logging procedures at the trench were used, with the

addition of trench photography to document the trench wall details. The trench face was
cleaned using trowels, and a leveled 1 m by 0.5 m reference coordinate system was affixed

to the wall with nails and twine. Before the trench was logged, map units were defined
based on lithology and sedimentary features. Contacts and sedimentary or structural
features were manually measured and plotted on a hand log at a scale of 1: 20. To
document details of the trench, including pebble and lamination location and orientation, a

photo-mosaic of the trench was created. A stationary camera frame was built that
maintained a 1 m distance from the trench wall, and a 24 mm wide-angle camera lens was

used to photograph the wall.


Computerized rectifications were applied to scanned photos of the trench to
correct for spherical aberrations from the camera lens, but this correction was later
abandoned because it significantly reduced the resolution and accuracy of the trench wall

details. Instead, information from each photo was traced and computerized to match with
other photos as accurately as possible. A hand-sketched log provided a framework for the
photo details in the preparation of the final trench log presented in this study.

GEOLOGIC SETTING

Tectonic Setting of the Himalaya


The Himalaya is divided into three major tectonic zones, bound by laterally

extensive, north-dipping faults (Gansser, 1964). From north to south, these faults are the
Main Central thrust (MCT), Main Boundary thrust (MBT), and the Himalayan Front

thrust (HFT) (Figs. 1, 2). At depth, the MBT and HFT probably merge into the Main
Himalayan thrust (MHT of Zhao, 1993), which is the decollement where large
earthquakes occur (Ni and Barazangi, 1984).
The northward-subducting Tethyan oceanic crust separated the Indian subcontinent and its passive northern margin from the Eurasian plate during the Late

Cretaceous through early Eocene (Brookfield, 1993). Initial contact of India with Eurasia

occurred as early as 65 Ma (Yeats and Hussain, 1987). By about 50 Ma, the Indian plate
reached the northern subduction zone, marking the beginning of the Himalayan collision

(Gansser, 1964; Le Fort, 1975; Patriat and Achache, 1984). Obducted Tethyan ophiolites
are preserved along the Indus-Tsangpo suture zone, north of the MCT (Molnar and
Tapponnier, 1975). Suturing of India and Eurasia was complete by 44 Ma, and uplift
subsequently began along thrust faults to the south (Patriat and Achache, 1984). The
remnant Tethyan ocean basin slowly filled in with shallow-water marine deposits and

eventually with brackish-water deposits (Najman et al., 1994) as the post-suture collision
culminated and Indentation tectonics began (Tapponnier and Molnar, 1977).
As much as 27% of overall shortening was accommodated by lateral translation of
Asian blocks north of the Himalaya by a rigid India converging from the south (Peltzer
and Saucier, 1996). Approximately 73 % of collisional shortening has been taken up by
thickening of the lithosphere along the MCT, MBT, and HFT (Peltzer and Saucier, 1996).
The MCT was active around 22-19 Ma, during maximum crustal shortening and uplift

(Hodges et al. 1992; Najman et al., 1994). The MCT separates crystalline rocks of the
Higher Himalaya from metasedimentary nappes within the Lesser Himalaya. The Lesser

Himalaya was thrust over Sub-Himalayan strata along the MBT, which was active from
about 11 Ma to 1.2 Ma (Valdiya, 1986; Burbank and Raynolds, 1988; Meigs et al., 1995).

The general trace of the MBT is marked by large salients and reentrants (Fig. 1). The
HFT, first documented by Nakata (1982), is the southernmost Himalayan thrust. It is
exposed discontinuously along the southern Sub-Himalaya.

Short-term convergence rates along the Himalaya are constrained by geodesy,

whereas longer-term rates are determined by geology. The collision rate at the Himalayan
front in Nepal, calculated from geodetic data, is 20.5 2 mm/yr (Bilham et al., 1998). In
Nepal, a similar Holocene shortening rate based on terrace abandonment patterns is 21

1.5 nun/yr (Lave and Avouac, 1998b). These rates account for about 30-35 % of the 58
4 nun/yr overall collision velocity of the Indian plate with Asia (Freymueller et al., 1996).

Finite-element models by Peltzer and Saucier (1996), based on slip rates of faults
from the Himalaya to Siberia, suggest that convergence rates at the Himalayan front

decrease from east to west. In India, geodetic data are sparse (Bilham et al., 1998).
Geologically-determined minimum rates of 14 2 mm/yr (Powers et al., 1998) and 11.9

3.1 mm/yr (Wesnousky et al., in prep.), indicate that convergence between the Indian
Shield and the Himalayan foothills is slower in India than in Nepal (Molnar, 1987; Yeats

and Thakur, 1998; Wesnousky et al., in prep.). Convergence rates along the Pakistan
Himalayan front may be as low as 9 mm/yr (Leathers, 1987; Baker et al., 1987; Pennock
et al., 1989).

Regional Earthquake History


Great Himalayan earthquakes are inferred to nucleate on a moderately-dipping
ramp of the MHT, below the surface trace of the MCT (Ni and Barazangi, 1984;

Baranowski, 1984; Lyon-Caen and Molnar, 1985; Molnar, 1990; Yeats and Thakur, 1998)
(Fig. 2). Fault rupture propagates southward, causing slip along a shallow-dipping flat
beneath the sub-Himalaya (Yeats and Thakur, 1998). Low modern seismicity on the flat
indicates that it is locked (Yeats and Thakur, 1998).

During the past century, there have been four major earthquakes which have
heavily damaged the Himalayan foothills (Fig. 3). Slip at the front is accommodated by
large active anticlines and synclinal dun valleys (Nossin, 1971) above a predominantly

blind HFT (Yeats et al., 1992). Surface rupture along the discontinuous HFT has not
been identified with any specific large historical earthquakes in the Himalaya (Nakata,

1989). A 900-year paucity of earthquakes prior to this last century makes it difficult to
calculate recurrence intervals.

The seismic gap theory postulates that the hazard associated with large
earthquakes increases with time since the last earthquake along a particular fault (Sykes,
1971; Kelleher et al., 1973). Along the Himalaya, there are three major seismic gaps (Fig.
3), although smaller earthquakes have occurred in these gaps (Khattri and Tyagi, 1983;

Khattri, 1987; Khattri et al., 1989; Yeats et al., 1992). The Kashmir Gap lies west of the
1905 earthquake rupture zone, the Central Gap lies between the 1905 and 1934

earthquake rupture zones, and the Assam Gap lies between the 1897 and 1950 earthquake

ruptures zones. Khattri (1987) does not indicate a gap between the 1934 and 1897
earthquake rupture zones.
The 1905 Kangra earthquake heavily damaged the Kangra and Dehra Dun
reentrants and killed about 19,000 people (Middlemiss, 1905; Quittmeyer and Jacob,
1979). Middlemiss (1910) documented damage and prepared a map of intensity zones.
Two zones of maximum intensity (> VIII, Modified Mercalli Scale - MMS) in the Kangra

and Dehra Dun reentrants were separated by a 100-km stretch of lower intensity across

the Nahan salient (Molnar, 1990) (Fig. 4). A lack of documented coseismic surface
rupture indicates that this earthquake probably ruptured a blind HFT, which contributed to
uplift of the Mohand anticline and the Doon Valley (Middlemiss, 1910; Yeats and Lillie,

1991; Gahalaut and Chander, 1992). The Janauri anticline of the Kangra Valley has
probably been uplifted in the same way (Middlemiss, 1910). Damage reports indicate that

the 1905 rupture zone was 120-300 km long (Quittmeyer and Jacob, 1979; Seeber and
Armbruster, 1981; Molnar, 1990). An average slip displacement of 5 - 7.5 m, and perhaps
as much as 10 m, occurred during the Kangra earthquake (Molnar, 1990; Bilham et al.,

10

34 _+
AD 25 "4

1885 /

11.1*

KG
Lahore

1*OlIN

1905
1803,

30 - +

Dehra Dun

CG
ao

Delhi

1833+

1934
rmation
26 - *
N
+
4.

04

74 E

78

82

86

90

94

Figure 3. Rupture zones of the great 1897, 1905, 1934, and 1950 Himalayan
earthquakes (shaded areas). Stars represent the approximate location of other
large historic earthquakes. Fault plane solutions are shown for moderate
earthquakes (5.5 S M 7) recorded between 1963 and 1984. Note that
seismicity isconcentrated north of the Himalayan deformation front trace.
Seismic gaps are located between the great earthquake rupture zones. KG =
Kashmir Gap, CG = Central Gap, AG = Assam Gap. Modified from Molnar
(1990) with additional information from Khattri (1987) and Yeats and
Thakur, (1998).

11

MAGNITUDE
SCALE

4.0 - 4.9
5.0 - 5.9
6.0 - 6.9

Figure 4. Meisoseismal regions of the 1905 Kangra and 1803


Mathura earthquakes. The Kangra earthquake had two zones of
maximum intensity in the Kangra (K) and Dehra Dun (DD)
reentrants, separated by a lower zone of intensity across the
Nahan salient (N). MBT = Main Boundary thrust. Also shown
are magnitudes of large earthquakes between 1914 and 1975.
The 1803 earthquake was not gap filling like the 1905 earthquake,
but the Dehra Dun reentrant was damaged. Modified after
Quittmeyer and Jacob (1979) with information from Oldham
(1882) and Middlemiss (1910).

12

1998). Using damage and possible fault dimension information, a M8

8.0 was assigned

to the Kangra earthquake by Kanamori (1977).


The 1803 Mathura earthquake was considered a smaller and non gap-filling

earthquake, although it damaged a wide area (Fig. 4) and was estimated at Ms 7-8, with
an epicenter about 80 km from Dehra Dun (Oldham, 1882; Quittmeyer and Jacob, 1979;
Mohindra and Thakur, 1998). Records of intensities indicate that most of the damage

occurred near Dehra Dun (VIII- IX, MMS) and also near Badrinath (VII- IX, MMS) to
the north. In the Sirmur district, which includes the Trans-Yamuna area, puka buildings
and mosques collapsed and several villages were destroyed (Oldham, 1882). For
clarification, puka (or pucca) buildings are made with stone or burnt brick in a mortar of

mud or lime (Arya, 1996). The 1803 earthquake may not have been large enough to
release the accumulated strain required to fill in the Central Seismic Gap (of Khattri, 1987)

between the 1905 and 1934 rupture zones (Bilham et al., 1998). However, it may have
been large enough to locally preserve seismically-deformed sediments.

Another large earthquake in 1833 struck the eastern part of the Central Seismic
Gap (Fig. 3), but it was probably not large enough to affect the Doon Valley area (Bilham,

1995). The epicenter, near Kathmandu, with an estimated moment magnitude of 7.5 < M
< 7.9, was too small to release accumulating elastic strain above the locked decollement in
that seismic gap (Bilham, 1995).

Stratigraphic Setting of the Himalayan Foothills


The Himalayan foothills of India include Precambrian strata of the northern Indian
passive margin, Tertiary shallow marine sediments of the closing ocean basin, and

freshwater molasse sediments of the rising Himalaya. A regional stratigraphic column for
the Lesser and Sub-Himalaya of the Nahan salient and Dehra Dun reentrant show these
relationships (Fig. 5).
Precambrian passive margin metasediments of the middle Proterozoic Shall
(Deoban/Tejam) Group includes slate and stromatolitic limestone (Srikantia and Sharma,

13

Figure 5. Regional stratigraphy of the Himalayan foothills of India.


General stratigraphic relationships and descriptions are from Auden (1934),
Karunakaran and Rao (1976), Thakur (1992), Brookfield (1993), and Najman
et al. (1993; 1997). Adapted from Powers et al. (1998).

14

LITHOLOGY

AGE

cc.,

o
.co

GROUP/ FM

DESCRIPTION

Doon
Gravels

conglomerate of plains
and duns; river terraces
and alluvial fan deposits

/
4)

c._

Upper
Siwalik
Formation

conglomerate
with increasing sandstone
away from MBT; some
calcite cementation; clay
and siltstone interbeds

Middle
Siwalik

sandstone
with minor claystone;
conglomerates prominent
closer to MBT; appearance
of kyanite as marker

,e .

Formation

Lower
Siwalik
Formation

Kasauli
Formation
(Upper Murree/
Dharamsala)

a)

o
o
0,

___ _

...

a
oI cp

Dagshai
Formation
(Lower Murree/
Dharamsala)

Subathu
Formation

W0

500+

of foothills

a)

a-

THICKNESS
(m)

2
0

2300

2000

alternations of sandstone
and claystone with
minor siltstone and
pebble horizons

1300

greenish grey
sandstone with minor
claystone

1300

purple clay and


siltstone with minor
sandstone

1300

red and green


nummulitic shale,
minor limestone
and sandstone

1500

limestone, siltstone,
shale and conglomerate,
in part weakly metamorphosec

2000+

CL c.)

""T""""..".",w"-"#:
E F3

--."

co-c
U
M

.0.-

---e. c

ca

.0
E
co

---

.,..--

....,..

Krol
Group

..0'.

_.....

gal,f3Kiarolini FInmfr:;Krol,

...,

,,--

----

,..--

......

--,1104S-1Jaunsar/Simla
-0-.
--- -,- - Group
-0,-- ...-. --

....es ...". ...0*".


...-- ...----

C.)
142

a.

(Nagthat, Chandpur,
Mandhall Fms)

2000+

limestone and slate

1000+

Shall Group
gejam/

Deoban)

Figure 5

quartzite, phyllite,
slate and marble

15

1976; Parkash et al., 1980; Thakur, 1992). Correlative to the Shali Limestone is the
Bilaspur Limestone of the peninsular Vindhyan Group, based on similar stromatolitic

layers (Karunakaran and Ranga Rao, 1976; Parkash et al., 1980). Unconformably
overlying the Shah Group is the upper Proterozoic Jaunsar/Simla Group, which includes
Mandhali slate and limestone, Chandpur phyllite and quartzite, and Nagthat quartzite

(Auden, 1934; Thakur, 1992). The Krol Group, as subdivided by Auden (1934), includes
conglomerate, siltstone, shale and limestone of the Blaini, Infra-Krol, and Krol formations

of uppermost Proterozoic and Lower Cambrian age (Auden, 1934). Thakur (1992) also
includes the Cambrian Tal Formation in the Krol Group.
The Sub-Himalayan foredeep, south of the MBT, comprises Tertiary sediments

that record the transition from marine to continental deposition during the early phases of
collision. Upper Paleocene to lower Eocene Subathu nurnmulitic shales were deposited on
an open shelf during the final basin closure (Najman et al., 1994). Upper Oligocene

28

Ma) to lower Miocene Dagshai and Kasauli clay and sandstone, temporally equivalent to

the Murree and Dharamsala Groups, are the oldest exposed continental foredeep deposits
(Najman et al., 1997). The middle Miocene Lower Siwalik Formation, in gradational
contact with the Kasauli Formation, is characterized by thick-bedded multistoried
sandstone, with mudstone and siltstone couplets. The upper Miocene Middle Siwalik
Formation comprises gray massive salt and pepper multi-storied sandstone, which grades

into Upper Siwalik conglomerate, with sandstone and minor mudstone. Quaternary
gravels of alluvial fans and post-Siwalik foredeep sediment record ongoing Himalayan
uplift and erosion (Meijerink, 1974).

Doon Valley Structure


The Doon Valley is a synclinal valley 80 km long and over 20 km wide. It is bound

to the south by the Siwalik range, uplifted along the HFT, and to the north by pre-Tertiary
rocks of the Lesser Himalaya (Fig. 6). The eastern closure of the Doon Valley is

16

3030'
GH

400+

Yamu
tear fault

IVIUSSO:0,-- rif,A

1/4/44s), DD.

4<s,
41(s,

Doon and Gangetic Plain sediments

Siwalik (Lower, Middle, Upper)


3000'

Undifferentiated Subathu and


Dagshai Fms with Shall fault slivers

Lesser Himalaya metasediment

0
77 E

20 km

Ganga
fault

dffi(t
78

Figure 6. Generalized geological map of the Doon Valley region, northwest

India. MCT = Main Central thrust, MBT = Main Boundary thrust, HFT =
Himalayan Front thrust, DD = Dehra Dun, GH = Garibnath Hill. Area
enclosed in box is shown in Plate 1 and Figures 7 and 8. Modified from
Powers et al. (1998).

17

controlled by the north-south Ganga tear fault (Raiverman et al., 1983; Arur, 1982), and
the western closure is partially controlled by the Yamuna tear fault (Valdiya, 1976).

Balanced structural cross sections constructed by Powers et al. (1998) show that
the Doon Valley has shortened by 11 km in the last 0.7-0.8 My, with 5 km accommodated
by the Mohand anticline. These data yield a maximum Doon Valley shortening rate of 15
1 mm/yr, although ages used in this calculation are not well constrained (Powers et al.,

1998). A Holocene horizontal convergence rate at the Doon Valley deformation front is
11.9 3.1 mm/yr (Wesnousky et al., in prep.). This rate is based on offset of uplifted
fluvial terraces near the HFT that are capped by a Holocene soil.

18

TRANS-YAMUNA GEOLOGY

The Trans-Yamuna area is located in the northwestern Doon Valley, at the


structural transition between an imbricate thrust fan of the Nahan salient and the broad
synclinal Dehra Dun reentrant (Fig. 6). Road access in the Trans-Yamuna area is limited
mainly to the Giri and Tons River valleys, and landslides often cause temporary closure of

these roads. More than 60 villages are located in the Trans-Yamuna countryside, thus

providing a network of footpaths in the populated areas. In more densely-forested parts


of the field area, footpaths are limited or absent.

Terrain of the Trans-Yamuna area varies in both relief and vegetation, as a

function of structure and lithology. Elevation increases from 480 m in the south to 1400
m in the area north of the MBT. Landsat TM satellite imagery at 30-m resolution clearly
depicts vegetation and topographic relief in the Trans-Yamuna area (Fig. 7). A dense
forest covers the southern and western areas, which is underlain by Siwalik rocks or postSiwalik fan deposits.

Local Bedrock Stratigraphy and Structure


The Trans-Yamuna area is cut by three major thrust faults, which are shown on the

geologic map included in this study (Plate 1). From north to south, these faults are the
Krol (MBT), Bilaspur, and Nahan thrusts. All three faults are north-dipping and southverging.

The upper Proterozoic to Lower Cambrian Jaunsar/Simla and Krol groups,


mapped as one undifferentiated map unit (Plate 1), crop out only north of the MBT

(locally named the Krol thrust). Dark gray to green resistant pyritic slate with uncommon
thin quartzite, chert, and limestone interbeds, exposed along the Girl River; is assigned to

the Mandhali Formation, by Vohra et al. (1976). At Milkan ka Khala, near Sataun, there
is a small section of green and purple slate, possibly of the Nagthat

30033,

Trans-Yamuna Active Fault System

3032'
N

7 7 40' E

77 45'

Figure 7. Landsat Thematic Mapper satellite image showing the surface expression of the
Trans-Yamuna active fault system, northwestern Doon Valley, India. Spatial resolution
is 30 m. Image processsed in the Remote Sensing Lab of the Indian Oil and Natural Gas
Corporation, Dehra Dun.

20

Formation. East of the Girl River area, massive, fine-grained, dark blue to gray limestone
and less resistant bedded and deformed shale (Krol Formation), both with quartz veins,

crop out in the northern part of the map area. Weakly-metamorphosed pebble
conglomerate of the Blaini Formation is exposed northeast of Dhamaun village. Slate
along the Tons River is part of the Mandhali Formation (Sinhval et al., 1973).

East of the Giri River, the Krol thrust juxtaposes rocks of the Jaunsar/Simla and
Krol groups against rocks of the Lower Tertiary Subathu and Dagshai formations (Plate).
Locally, the Krol thrust brings Jaunsar/Simla rocks directly over Miocene Lower Siwalik,

similar to relations in the north central Doon Valley (Raiverman et al., 1983). Along the
Giri River, in the western part of the map area, the Krol thrust brings rocks of the
Jaunsar/Simla and Krol Groups over older Precambrian rocks of the Shalt Group.
The MBT is generally defined as a thrust that separates the pre-Tertiary from the

Tertiary. Local tectonic slivers of Precambrian rocks within the Tertiary rocks, however,
complicate this tectonic relationship. Imbricate faulting in the Nahan salient to the west
exposes limestone of the Shaft Group within thrust wedges of Lower Tertiary Subathu and
Dagshai formations (Auden, 1948; Bhargava, 1976; Srikantia and Sharma, 1976; Vohra et

al., 1976). Because Lower Tertiary rocks are found north of the Shall thrust wedges
within the Nahan salient, the MBT has been placed north of these thrust wedges
(Karunakaran and Ranga Rao, 1976; Raiverman et al., 1976; 1981; 1983).
The Shalt Group is older than the overthrusted Jaunsar/Simla and Krol groups
along the MBT (Fig. 5). Large discontinuous lenses of Shaft Limestone (also locally

referred to as Bilaspur Formation or Sataun Formation) are gray and stromatolitic, with
smaller black deformed lenses of shale and siltstone. In some localities within the Nahan
salient, lower Tertiary rocks unconformably overlie the Shaft within the imbricate fan

(Srikantia and Sharma, 1976).


Within the Trans-Yamuna area, Subathu and Dagshai formations are not found
north of exposures of the Shaft Group, but a distinction is made between the Krol thrust

(MBT) to the north and Bilaspur thrust to the south of the tectonic wedge. Isolated Shalt
outcrops within the Sub-Himalaya are referred to as the Shah subsidiary belt (Srikantia

and Sharma, 1976). The Shah Group does not crop out east of Bhatrog (Plate 1).

21

Termination of the Bilaspur thrust in the western part of the Trans-Yamuna area is

probably related to eastward lensing out of the broad Nahan salient at the western Dehra
Dun reentrant.
Lower Tertiary Subathu and Dagshai formations are grouped as a single map unit

on the Trans-Yamuna geologic map (Plate 1). The Subathu Formation, limited to the
Sataun area, is a dark green to red, thinly bedded argillite with rare limestone interbeds.
The Dagshai Formation is a blue-gray resistant muscovite-rich sandstone with red
clay/shale sheared interbeds. The Kasauli Formation has not been found in the TransYamuna area.

The outcrop belt of lower Tertiary rocks is wide west of Sataun and narrower east
of Sataun. At the Tons River, their distribution widens again, but they are entirely absent
along the northern Doon Valley, east of the Trans-Yamuna area.

The Nahan thrust is a north-dipping thrust south of the MBT, with Lower Tertiary
rocks on the hangingwall and Upper Tertiary Siwalik rocks on the footwall. There are
nomenclature disagreements associated with the Nahan thrust, but this fault relationship is
the same as that described by Auden (1934), Sinvhal et al. (1973), Ansari et al. (1976),

Vohra et al. (1976), Rao (1978; 1986), Valdiya (1980; 1992), and Swamy (1986). In the
Trans-Yamuna area, the Lower Siwalik Formation crops south of the Nahan thrust. This
formation comprises a light-brown well-sorted, medium-grained sandstone with planar
laminations and low-angle cross bedding. It is locally interbedded with a reddish-brown

claystone. Sandstone of the Lower Siwalik Formation is less resistant to weathering than
sandstone of the Dagshai Formation.

The Nahan thrust is continuous and separate from the other bedrock thrusts at the
surface in the western part of the study area (Plate 1). The Nahan and Krol thrusts in the
east maintain less well-defined surface traces, indicating a different fault dip geometry.

The intersecting branch line between these two faults is near the surface in the eastern part
of the map area. This expression is reflected by small discontinuous slivers of the Dagshai

Formation (see Plate 1 and cross section D-D' therein). The Nahan thrust may dip more
steeply than the Krol thrust, as inferred by outcrop patterns of rocks bounded by the

22

faults. The Nahan thrust is discontinuous east of Sataun, and it disappears where the
Nahan and Krol thrusts meet or where one thrust is cut by the other.
The Krol thrust trace near Nau village changes in strike from east-west to north-

south orientation with no field evidence for a transfer fault. At depth, the Krol thrust may
form a lateral ramp such that the hangingwall geometry results in an apparent left step.
This structural style is similar to that of the Gillibrand Canyon lateral ramp of the Santa

Susana fault in California (Yeats, 1987) and the Lewis thrust in Montana (Boyer and
Elliot, 1982).
A Tertiary microgabbro to gabbro dike cuts discontinuously through the western
part of the field area with a northwest-southeast strike (Plate 1). It is a 0.5 to 10 m-thick
coarse-grained gabbro dike within the Shah Formation, west of Manal, near the Krol

thrust. At Sataun, a 1 to 15 m-thick microgabbro crops out along the Bilaspur thrust and
within the Subathu and Dagshai formations. Just east of Sataun, the dike crops out north
of the Bilaspur thrust and intrudes the Shah formation. This is the eastern extent of the

dike outcrop. This dike, not restricted to fault zones, was previously recognized at Sataun
and assigned a probable post-Eocene age by Vohra et al. (1976).

A north-south striking fault near Sataun is transverse to the structural grain. This
fault, referred to in this study as the Malgi fault, is expressed as a strong lineament in

satellite imagery and on topographic maps. This fault offsets the Nahan thrust south of
Sataun. It is straight and stepped en echelon to the left, similar to strike-slip faults
documented approximately 15 km to the west by Swamy (1986).

Quaternary Deposits and Geomorphology


Pleistocene to Holocene alluvial fan deposits correlative to the Doon Gravels
(Nossin, 1971; Meijerink, 1974; Rao, 1977; Khan and Dubey, 1981) cover the Trans-

Yamuna slopes. The proximal-fan facies of the steeper slopes of the Trans-Yamuna area
are commonly debris-flow deposits. These deposits are poorly sorted with angular clasts
of pre-Tertiary Lesser Himalayan metasediments in a light brown silt matrix. Mid- to

23

distal-fan facies comprise moderately sorted, weakly imbricated, subrounded pebbles and

cobbles with sand, and dip approximately 10 to the south. Locally, mid-fan deposits are
calcified (Nossin, 1971) and form cliffs as high as 20 m.

The older, large alluvial fan in the Trans-Yamuna area extends southward beyond

the study area. Garibnath Hill, about 7 km south of the study area (Fig. 6), is a fan
remnant that suggests post-depositional downcutting of approximately 137 to 165 m by

the Giri River (Nossin, 1971). A fluvial terrace deposit at the northern end of Garibnath
Hill is the highest fluvial deposit above the Giri River base-level (Nossin, 1971). There are
three lower terrace levels within the Trans-Yamuna area (Plate 1 and Figure 8).

24

7735' Topographic base from the Government of India (1970) Sheet 53 F/10 & F/14 (1:50,000)

Explanation

Active Fault- Drawn where scarps present. Amount and direction of dip shown where known.
Ball and bar shown on downthrown side of fault
Fault- Dashed where approximately located; dotted where concealed. Amount and direction of dip
shown where known
Thrust Fault- Dashed where approximately located; dotted where concealed, sawteetll on upper
plate. Amount and direction of dip shown where known
Fluvial Terrace Deposit (Holocene and Pleistocene)- Terrace numbers represent relative height levels
above river base. Qt3 is the highest (35-80 m), Qt2 is mid-level (15-35 in), and Qti is the lowest (2-15 m).

Figure 8. Topographic map of the Trans-Yamuna area, with faults and fluvial terraces.
Terrace group assignments modeled after Nossin (1971) and Khan and Dubey (1981).
KT = Krol thrust, BT = Bilaspur thrust, NT = Nahan thrust, MF = Malgi Fault,
SF = Sirmurital fault, DF = Dhamaun fault, BF = Bharli fault. See Plate 1 for corresponding
geology of the Trans-Yamuna area.

25

TRANS-YAMUNA ACTIVE FAULT SYSTEM

The Trans-Yamuna active fault system (TYAFS) is a left-stepping en echelon

group of faults approximately 14 to 19 km long. Fault lineaments and general slope


breaks associated with the Trans-Yamuna area were documented initially by Nossin

(1971) and Nakata (1972) as part of larger regional studies in the Doon Valley. Detailed
analysis of digital satellite imagery at 30 m resolution (Fig. 7), coupled with geologic

mapping and associated field work, shows that these faults cut Quaternary deposits of
probable late Holocene age.
The TYAFS is subparallel to and crosses the Krol thrust (MBT) with a distinctive
south-side-up trace, similar to active faults along the Nepal MBT (Nakata, 1982, 1989;

Nakata and others, 1984). On satellite images, these faults have strong shadows with
illuminated traces that are opposite the bedrock thrust faults. This indicates that the north

side is up (Fig. 7). The MBT does not appear to be active in the Trans-Yamuna area. The
TYAFS may have, however, followed a plane of weakness formed by the MBT, especially

where the MBT formed a schuppen zone with secondary thrust faults in the Sataun area
(Plate 1).

The topographic and geologic maps show the TYAFS in red with a ball and bar on

the downthrown side (Fig. 8 and Plate 1). Dashes along the active faults show the stepped
pattern and do not indicate location uncertainty as they do for the inactive faults. The dots

that are used to show concealed inactive faults are not used for the active faults. The
active faults are described where they are geomorphically expressed at the surface. The
TYAFS includes three major subdivisions, referred to here as the Sirmurital, Dhamaun,

and Bharli faults. Each mapped and described fault is separated by surficial overlaps or
fault trend shifts.

26

Sirmurital Fault
The Sirmurital fault has an east-west, slightly convex northward trace over its

approximately 4.5 km length in the western part of the field area. It is expressed clearly
both on satellite imagery (Fig. 7) and on a topographic map (Fig. 8), west of the Giri

River. The Sirmurital fault cuts Lower Siwalik rocks, which are poorly exposed due to
cover by a thick jungle canopy. The approximate fault scarp height within Siwalik rocks is

20 m. Also within Lower Siwalik rocks, the Sirmurital fault appears to cut the Malgi
fault. The Malgi fault is left-stepped, however, so apparent fault separation caused by the
Sirmurital fault may be exaggerated (Plate 1). The Sirmurital fault extends eastward
through Quaternary fluvial terraces on the west side of the Girl River, near Sirmurital

village (Fig. 9). The apparent scarp height of the Sirmurital fault in the terrace area is
approximately 55-60 m, significantly higher than other scarp heights along fault. There
may be an interaction between fault displacement and deposition of fluvial terraces, such

that the scarp height reflects both fault separation and the natural terrace riser between Qti
and Qt3. With an overall scarp height of 20 m or more, faulting reflects both Holocene
and Pleistocene motion.

Dhamaun Fault
The Dhamaun fault oversteps the Sirmurital fault by approximately 1 km (Plate 1).

Modern alluvium of the Giri River obscures the relationship of the Dhamaun fault with the

eastern terminus of the Sirmurital fault, and these faults may connect at depth. The
Dhamaun fault trace is characterized by a series of left-stepping 5-10 m high scarplets over

its approximately 6.5 km length. Near Sataun, the Dhamaun fault is parallel to the trend
of the Krol, Bilaspur, and Nahan thrusts, all of which lie within 1 km of each other. The
Dhamaun fault extends east through steep topography and cuts the Krol thrust (MBT).
The Dhamaun fault cuts the upper terrace (Qt2) near Bhatrog village, on the east
side of the Giri River (Fig. 10). The apparent fault motion is reverse with a 57 north dip,

Figure 9. Photograph of Sirmurital terraces (far side) and Bhatrog terraces (near side) of the Giri
River. Upper (Qt3) and lower (Qt i) Sirmurital terraces are 70 and 12 m above the Giri River, respectively.
Upper (Qt2) and lower (Qt Bhatrog terraces are 18 and 2 m above the Giri River, respectively. The
Sirmurital fault strikes between the upper and lower Sirmurital terraces. Arrows indicate other figure
locations along the Giri River. The high scarp at the arrow marking Figure 15 reflects both the fault offset
and the terrace riser of Qti

28

Figure 10. Photograph of the Dhamaun fault (yellow arrows) exposed along
the Giri River. View is east toward the upper Bhatrog terrace (Qt2), which is
in fault contact with the lower Tertiary Dagshai Formation (LT). The fault
dip is 57" north, and apparent fault motion is reverse despite a geomorphic
south-side-up trace elsewhere along the fault system. See Figure 9 for
location. The person standing on the fault trace is 1.6 m tall.

29

despite the dominant geomorphic south-side up expression elsewhere along the trace.
Dagshai sandstone and claystone are in the hangingwall, with gravel that overlies Dagshai
claystone in the footwall.

On the west side of Giri River, about 200 m north of the Sirmurital fault, terrace
pebbles are entrained in clays of the Dagshai claystone along the Dhamaun fault (Fig. 11).

The claystone at this site is highly sheared with an east-west strike and 55-65 south dip of
the shear plane.

Farther to the west along the south side of Gin River, near the northern
termination of the Malgi fault, the Dhamaun fault brings the Dagshai Formation over

fluvial terrace gravels deposited on the Dagshai Formation in the footwall. The fault
relationship here is similar to that described at the upper Bhatrog terrace, approximately

1.5 km to the east, except that the apparent fault dip is steep to the southwest, with the
south side up.

A 2 to 4 m ridge within bedrock, north of the lower Sirmurital terrace surface, is

sub-parallel to the west terminus of Dhamaun fault (Plate 1). The ridge extends south
beyond bedrock, along the surface of the lower Sirmurital terrace, where it comprises

terrace strata. Streams developed on the lower terrace are deflected around the bedrock
portion of the ridge, leaving an erosional terrace remnant from stream incision. This is the
southern expression of the topographic ridge. There is no evidence that this ridge is a
structural "pressure ridge" similar to that described by Nakata et al. (1984).

Bharli Fault
The Bharli fault strikes east-west for about 4 km, east of the Dhamaun fault (Plate
1). Its linear expression suggests strike-slip motion, although evidence for strike-slip is

limited to two prominent sag ponds southeast of Bharli village, both of which could be
due simply to the south-side-up nature of faulting in a high relief area (Fig. 12). High
resolution satellite imagery suggests left-lateral offset of a stream along the Bharli fault,

30

Figure 11. Photograph of entrained Giri River terrace pebbles (p) in sheared
Dagshai Formation along the Dhamaun fault. View is west. Sheared material
dips 55-65 south. See figure 9 for location.

Figure 12. Photograph of a sag pond on the north side of Bharli fault. View is to the southwest. The
fault scarp here is 8-10 m high.

32

near the eastern sagpond, but the field relations of the stream channel cannot be
established because of fan and talus cover.

North of Rajpur Village, the Bharli fault has a scarp height that ranges from 3 to 7
m, with the same distinctive south-side-up trace throughout the TYAFS (Plate 1). It cuts
the sinuous Krol thrust (MBT) at a local 30 m left step just east of Rajpur village, similar
in style to the western site where the Dhamaun fault cuts the Krol thrust, as previously

noted. The geological map shows a fault separation where the Krol thrust is cut by the
Bharli fault, due to south-side-up faulting (Plate 1).

Farther east, the Bharli fault is exposed just south of the eastern sag pond, on the
west side of the road leading to Bharli Village (Fig. 13). The fault dips 54 to the south,
with an apparent reverse motion. Krol phyllite exposed on either side of the fault is
heavily sheared and deformed. Fan deposits are not offset by the fault at this site,
although the fault trace continues laterally into fan deposits with scarps. There is no
evidence for strike-slip motion at this exposure.

The eastern termination of the Bharli fault may be related to pre-existing bedrock
structure such as the possible lateral ramp of the Krol thrust near Nau village (Plate 1)

discussed previously. The Krol thrust ramp probably hinders propagation of the Bharli
fault further east.

Other Related Active Faults


There are additional active faults in the eastern Trans-Yamuna area, and some may

be related to the TYAFS. Three small faults south of the Bharli fault (Plate 1) are linear
with a south-side-up displacement. The fault just west of Nau village has a clear
geomorphic expression with uplifted calcified fan deposits on the south side of the fault.

A small fault south of Tauru has a 2 to 3 m scarp with a small sag pond on the north side.
West of Bobri village, a strong geomorphically expressed lineament marks a fault that

continues into fan deposits at both ends.

33

Figure 13. Photographs of a Bharli fault exposure. This site is 0.5


km west of that shown in Figure 12. a) View west of south-side-up
faulting. Box shows location of area in lower photograph. b) Closeup of the fault exposure (yellow arrows), here within bedrock.
Dip is 54 to the south.

34

The fault at Kalawar village (Krishnaswamy et al., 1970; Nossin, 1971; Sinvhal et

al., 1973) cuts fluvial terrace deposits, but has the south-side-down, which is the opposite
of other active faults in the Trans-Yamuna area (Fig. 14 and Plate 1). This fault appears to

be a local reactivation of the Nahan thrust, based on the location of the fault and bedrock
relationships near the Tons River. An apparent 40 m right-lateral offset of the two
terraces at Kalawar indicates possible strike-slip or oblique-slip motion. If the fault were
active before and after deposition of the lower terrace, vertical displacement on the scarp

at the upper terrace should be greater. Instead, the scarp cutting the upper terrace (14 m)
is 6 m less than the scarp on the lower terrace (Fig. 14), perhaps because the fault dies out
to the west or it slips obliquely. If this fault is related to the TYAFS, than the system
length would be as much as 19 km.

35

Figure 14. Photograph and simplified sketch of an active fault at Kalawar


Village with 40 m of apparent right-lateral separation. This east-west fault
cuts terraces with the north side up rather than the prevalent south-side-up
expression of the Trans-Yamuna active fault system. The fault shown here
may be a local reactivation of the Nahan thrust. Yellow arrows in the photo
show the trace of the fault where it cuts Qt l and Qt2 terraces of the Tons
River. The fault, with balls and barbs on the downthrown side, is shown in
red. The black bars with hachures represent the terrace riser between Qti
and Qt2. The numbers indicate fault separation and terrace riser heights.
The view of this photo is to the south, overlooking the Yamuna valley.

36

PALEOSEISMIC INVESTIGATION AT SIRMURITAL

The purpose of paleoseismic investigations is to document deformation within

sediments related to a fault that is presumed to have ruptured in the late Quaternary, in

order to establish a prehistoric record of large earthquakes. Sediment dating by


radiocarbon, thermoluminescence, or soil correlations helps to constrain prehistoric
earthquake timing to determine if a fault is active. Paleoseismic investigations commonly
involve an excavation at or near a fault to look at the near-surface sediment deformation.

Site Description

Factors involved in the choice of a site for a paleoseismic investigation included

topographic relief, access, scarp height, location of faulted Quaternary sediments, and the

type of sediments involved. The Himalayan foothills are steep, the roads are sparse, and
the village density is high. These circumstances restricted the choices of potential trench
sites.

There are many locations along the Trans-Yamuna active-fault system where
Quaternary sediment is cut, but only a few where depositional conditions could yield
earthquake information. A site along the Bharli fault was chosen where sediments
reflective of sag pond damming and fluvial deposition intervals could potentially indicate

earthquake history. Unfortunately, sedimentary relations in that excavation were obscured


by the coarse nature of the slope wash and fan material. Another site at Sirmurital was
more favorable, despite a large apparent scarp height.
The Sirmurital fault extends between two easily accessible terraces (Qti and Qt3)

near the Gin River, the lower of which comprises fine-grained deposits, with organic

material for radiocarbon dating. An apparent scarp height at Sirmurital of 55 to 60 m,


which is notably larger than other 10 to 20 m scarps along the TYAFS, is probably due to

faulting with respect to the fluvial terrace riser. Most terraces along the Gin River are

37

naturally stepped downstream, but at Sirmurital, the downstep is upstream (Fig. 7). The
trench was excavated into the lower Sirmurital terrace, near the fault scarp. Fine-grained,
north dipping strata exposed in the trench were logged and photographed (Fig. 15).

Trench Stratigraphy
The trench wall was divided into eleven map units for logging, which were

differentiated on the basis of lithology (Plate 2). Unit 1, the stratigraphically lowest,
comprises predominantly intrabasinal angular to subangular sandstone cobbles and

boulders of the Lower Siwalik Formation in a red silt matrix. Some angular quartzite
pebbles are also present in this unit. Unit 2 comprises well-sorted sand with pebbles,

fining upward to fine-grained sand, with minor charcoal present. Unit 3 is also sand, but is
coarse- to medium-grained with one cross bed

25 cm). Unit 4 is a purple to gray sticky

clay with fine laminations and sand lenses. Unit 5 is similar to 4 but more blue in color

and with some internal deformation. It also grades laterally to a fine sand near its
southward pinchout. Unit 6 comprises thin peat and medium-grained sand interbeds,

which grade laterally to sand laminations. Above grid section 5-6D (Plate 2), sand in Unit
6 appears to load and deform the underlying clay units. Unit 7 comprises poorly-sorted
angular intrabasinal sandstone pebbles, cobbles and boulders with minor metamorphic

pebbles. Unit 7 pinches out to the north such that deposition between units 6 and 8 is
continuous. Unit 8 is similar lithologically to unit 6, except that the peat and sand
interbeds are slightly thicker. Unit 9 comprises poorly-sorted laminated medium-grained

sand with abundant charcoal and pebbles of sandstone and quartzite. Unit 10 also
comprises poorly sorted medium-grained sand, but with an increasing amount of
metamorphic pebbles and some internal deformation. The amount of cobbles and boulders

within unit 10 increases to the south. Unit 11 is a thick, uniform deposit of poorly-sorted,
subrounded to rounded extrabasinal metamorphic cobbles and boulders with some
sandstone in a matrix of sand and pebbles.

38

Figure 15. Photograph of Sirmurital trench, showing stationary


camera frame and shoring device. Trench grid is 1 m wide by 0.5 m
high. View is southwest. See figure 9 for location.

39

Units 2 and 3 of the lower Sirmurital terrace may be reflective of either a meander
cutoff or overbank deposition. Only one crossbed is present in the mostly laminated, fine-

grained sand deposits of units 2 and 3 (Plate 2). Given the basin geometry and possible
flow direction, these sands are probably not point-bar deposits. This is supported by
southward-tapering (albeit north-dipping) sediments of most map units within the trench,

perpendicular to the flow direction. The north-tapering exceptions are units 1 and 7.
These two deposits are probably localized mass-wasting events, indicated by poor sorting,

clast angularity, and fine-grained matrix support of clasts. At some point, the basin at
Sirmurital was closed, which is revealed by gray clay deposits of units 4 and 5. Units 6
and 8 mark a gradual resumed influx of fluvial sands (annual flood events ?), interbedded

with marsh-like peat layers. Units 9 and 10 mark a return to a possible fluvial
environment, as suggested by laminated sands with some pebbles. Unit 11 is probably a
large flood deposit, similar to 2-3 m thick flood deposits documented in California
(Rockwell et al., 1984).

Radiocarbon Dating
Charcoal samples from Sirmurital were sent to two separate labs for radiocarbon
dating. Three samples were sent to Beta Analytic in Miami, Florida, for dating using

accelerator mass spectroscopy (AMS). Five samples were sent to the Birbal Sahni
Institute of Paleobotany (BSIP), Lucknow, Uttar Pradesh, India, for conventional
radiocarbon dating. Below is a table which summarizes the reported radiocarbon date
results for Sirmurital trench (see Plate 2 for sample locations). The reported ages, in
radiocarbon years, were converted to calendar years before 1950 using a tree ring
calibration to account for fluctuations of14C due to magnetic shield strength variations
(Stuiver and Reimer, 1993).

40

Radiocarbon ages for Sirmurital trench


Trench
unit

14C age

Calendar years
(Cal B.P.) t

34

10
8
8
8
8
7

20

1=

'modern'
310 40
610 100
1310 930
540 40
4820 320
1350 120
420 50

> 300 ??
474-287
711-486
3374-60
636-601/ 560-508
6284-4818/ 4755-4722/ 4669-4653
1506-982
531-315

Sample

number
8

7*
6
5

4*

* ba - Beta Analytic, Miami, FL, USA (AMS method);


bsi - Birbal Sahni Institute of Paleobotany, Lucknow, UP, India (conventional method)
t Cal B.P. (1950), Ca lib 3.03.c (Stuiver and Reimer, 1993),
results reported for a 2 sigma, 95% probability range

The sample ages are scattered with respect to stratigraphic position, but all

indicate that the lower Sirmurital terrace (Qtr) is of late Holocene age. This scatter may
be due either to sample contamination or detrital deposition. If samples 3 and 5 are
thrown out because of the large errors, the age sequence indicates deposition within the
last 2,000 years. If sample 2 is detrital, then the sequence of ages (samples 1, 4, 6, 7, 8)
decrease with the upward stratigraphic position, within the error bars, and suggests
deposition may have occurred in the last 600 years.

Quaternary Synclinal Folding at Sirmurital


A fault is not exposed in the Sirmurital trench, but north-dipping strata within the

trench are part of a Quaternary syncline that is parallel to the fault scarp. The trench
sediments dip south (Plate 2 and Fig. 15), whereas other exposures within the terrace that
are less than 100 m away from the trench dip the opposite way (Fig. 16). These two

exposures make up the north flank of the syncline at Sirmurital. One is an exposure along
the cut bank of the Gin River (Fig. 16a). It comprises 5 to 10 south-dipping fine-grained
sediment similar to that found in the trench, except without an updip sediment pinchout.

41

Figure 16. Photographs of 5 to 15 south-dipping strata within the lower


Sirmurital terrace, which form the north limb of a syncline that is parallel to
Sirmurital fault a) Exposure along Giri River with fine-grained sediment
similar to that of the trench. View is west. b) Exposure within stream cut
of the lower terrace. View is northwest. The person in the photo is 1.8 m tall.

43

The second exposure is located along a stream within the terrace, near the main Giri

channel (Fig. 9). Gravels at this site dip up to 15 southwest (Fig. 16b), and they also do
not taper updip. Folding appears to also affect the bedrock near the Holocene sediments
with the same orientation. There is no evidence that pre-existing topography affected the
dips on the north flank of the fold, where fine-grained deposits dip to the south (Fig. 16).

44

DISCUSSION

Style of the Trans-Yamuna Active Fault System

The TYAFS maintains a distinctive south-side-up trace throughout its length (Fig.

7). A linear fault trace suggests high-angle faulting. The three separately-described faults
may connect in the subsurface, although no ONGC seismic lines cross the MBT in this

area to test this idea. Fault exposures along the three faults described in this study reveal
surface variations which are due to pre-existing structure and lithology. Given the fault
traces, geometry, and limited exposures, the TYAFS is probably a high angle reverse fault
with oblique slip within the Sub-Himalayan hangingwall. It deforms to accommodate
uplift during deeper thrusting along the Main Himalayan fault.

The Sirmurital fault exhibits dip separation, but it is not clearly exposed anywhere

along its length. The eastern terminus of the Sirmurital fault is presumably at Giri River,

but the exposure of the fault there is obscured by scarp and terrace erosion as well as the
modern Giri River channel.

The Dhamaun fault is exposed at three locations along the Giri River, most clearly

near Bhatrog Village (Fig. 10). The north-dipping fault exposure and apparent reverse
motion are in contrast to the more common south-side-up expression. Its expression is

opposite that of another exposure at the west end of the Dhamaun fault. A component of
strike-slip motion can explain the near-surface dip and opposite separations.
Sag ponds at Bharli fault could either be related to simple valley-side up fault

ponding or it could indicate strike-slip motion. A few small streams appear to be laterally
offset, but the high relief and modern fan deposit mantle preclude accurate measurements
relevant to fault motion.

Faults with a length of about 18 km and 1 m surface deformation are capable of


Mw = 7 earthquakes (Wells and Coppersmith, 1994; McCalpin and Nelson, 1996). The
TYAFS length may be as great as 19 km if the other active faults in the Trans-Yamuna

area are related at depth to the TYAFS. Although the TYAFS is probably not long

45

enough to produce a great earthquake (Mw > 8), it is capable of producing secondary
faulting and deformation in the hangingwall as a secondary offset of large earthquakes on
the decollement (Wells and Coppersmith, 1994; McCalpin and Nelson, 1996).

Terrace Ages and Folding as Indicators of Fault Timing


It is important to determine the ages of sediments cut or deformed by the TYAFS

in order to establish timing of recent faulting. A Holocene earthquake history can be


worked out if a fault cuts datable sediments that can be matched up on both sides of the
fault. At Sirmurital, scarp erosion may have precluded finding the fault near the surface.
Paleoseismic trenching was useful, however, in that it provided fold information that may

be related to recent faulting.


A relative age estimation of the Bhatrog terrace (Qt2) that is cut by the Dhamaun

fault gives a first order age of faulting. An earthquake history cannot be established at this
site, however, because Holocene sediments are only preserved on one side of the fault

(Fig. 10). To determine the age of this terrace, an assumption is made that the
downcutting rate is the same for terraces of the Gin River valley over a distance of 7 km.
The fluvial deposit at Garibnath Hill, just south of the field area (Fig. 6), is at least 137 m
above the Giri River. If fluvial terraces within the Himalayan foothills are post-glacial

(Nakata et al., 1984; Lave and Avouac, 1998a), then a maximum age of 20,000 years can

be estimated for the terrace at Garibnath hill. If this deposit represents the maximum
height of Gin River, an estimated age for the upper Bhatrog terrace (19 m above the
modern river) is 2.7 ka. This indicates that the TYAFS has late Holocene movement.
Late Holocene synclinally folded sediment in the terrace at Sirmurital may indicate

that faulting along the TYAFS could be younger than the -3 ka age of the Bhatrog
terrace. The fold, exposed in the trench, can be explained by depositional conditions,
tectonics, or a combination of both. During deposition, sediment may have draped over
pre-existing topography, such that the basin was filled in with naturally-inclined sediments

and a decreasing dip up-section. A tectonic explanation for folding at the lower Sirmurital

46
is based upon very steep internal lamination dips of both sand and clay in the trench, lack
of evidence for pre-existing topography within the north flank of the syncline, and plunge

of the syncline away from the Gin River. A literature review by Thomas et al. (1987)
indicates that naturally inclined strata have dips ranging from 0 to 25, and that it is rare to
find naturally inclined strata at angles higher than 29 (Thomas et al., 1987; Miall, 1996).
Whereas this dip range is used for sand, laminations within the clay units of units 4 and 5

are also at high angles. The south-dipping sediments of the north flank (Fig. 16) neither

pinch out nor show evidence of drape. The consistent sediment dip there, opposite to that
found in the trench, suggests a tectonic influence. Within the trench section, an upward
decrease in dip suggests that folding may have been syndepositional.

Unit 7 (and perhaps unit 1) within the trench may be an earthquake-induced

colluvial wedge. Unit 7 exhibits a wedge shape that pinches out away from the scarp, and
most clasts within this unit comprise intrabasinal Siwalik sandstone. Units 2 through 6 are
steeper as a package than units 8 through 10, which onlap unit 7 at a more shallow angle.

The only recorded large earthquakes that damaged the Trans-Yamuna area, however, are
the 1803 and 1905 earthquakes. Radiocarbon ages for the sediments above the colluvial
wedge are older than 150 years, and the units below unit 7 are still with a historic time

range. If unit 7 was due to an earthquake prior to 1803, it is an unrecorded earthquake.


Units 8 through 10 may be tilted as a result of the 1803 or 1905 earthquakes, after a
previous event which may be attributed to unit 7. The lower terrace at Sirmurital may
record evidence of past earthquakes, which could have contributed to a possible broken
fault-propagation fold, such that the fault-parallel syncline is in the footwall.

Influence of Oblique Convergence Between India and Eurasia


Oblique convergence documented at oceanic subduction zones exhibit arc-parallel
strike-slip faulting, which may apply to Indian-Eurasian convergence (McCaffrey, 1992;

McCaffrey and Nabelek, 1998; Fig. 17). McCaffrey and Nabelek (1998) demonstrate that
although thrust slip vectors indicate arc-perpendicular motion of the downgoing Indian

47

Figure 17. Convergence diagram of India and north Tibet, showing


arc-parallel convergence south of the Karakoram fault. Arcuate
stippled line represents the Himalayan deformation front. Large
black arrows illustrate oblique convergence vectors between India
and north Tibet, calculated by McCaffrey and Nabelek (1998).
Smaller black arrows are thrust earthquake slip vectors. KF =
Karakoram fault, KFB = Karakoram fault block, TY = location of
the Trans-Yamuna area. See text for discussion. Modified after
McCaffrey and Nabelek (1998).

48

slab, convergence strain is partitioned between northern Tibet and India, such that arcparallel fault displacements have developed. The Karakorum fault primarily partitions the
strain, and the area south of this fault might be considered a block extruding to the
northwest at a rate of 20 mm/yr relative to northern Tibet (McCaffrey and Nabelek,
1998). Right-lateral strike-slip on the Karakorum fault may further explain a westward
decrease in convergence rates from Nepal to India (Yeats and Thakur, 1998).

If geodetic 20 mm/yr convergence rates are constant across the arc, then there
must be a slip deficit to account for lower geologically determined rates (Jackson and
Bilham, 1994; Bilham et al., 1997; 1998). Slip partitioning involving the Karakorum fault

could account for this discrepancy (Peltzer and Saucier, 1996), or it could be taken up by
a series of smaller out-of-sequence thrust faults within the western Himalaya (Yeats and
Thakur, 1998). The TYAFS, located within the Karakorum fault block (Fig. 17), may be
a response to such strain partitioning, such that the possibility of related strike-slip exists.
If this is so, then the TYAFS, and perhaps other similar faults north of the HFT contribute

to shortening rates that are unaccounted for by thrusting entirely by the HFT or by a
constant arc-wide convergence rate.

Large Magnitude Earthquake Recurrence and Seismic Hazard


The TYAFS is a relatively small fault system, but along with other active faults yet

to be documented, it may supplement low convergence rates in the western Himalaya to

account for a geodetic slip deficit (Bilham et al., 1998). Such faults may release a portion
of strain from hangingwall earthquakes. Higher potential estimates of convergence rates
in the western Himalaya have important implications for shorter recurrence intervals and

the associated seismic hazard in the Himalayan foothills. With an exploding population in
the Himalayan foothills, a large-magnitude decollement earthquake will cause tremendous

damage and loss of life. The absence of pre-historic earthquake makes it necessary for
continuing paleoseismic studies within the Himalaya, so that recurrence intervals can be

49

more tightly constrained. This is important for planning purposes and hazard mitigation in
northern India.

50

CONCLUSIONS

Satellite data analysis, geologic mapping, and paleoseismic trenching in the


northwest Doon Valley of India lead to the following conclusions:
1.

East-west lineaments identified in satellite images of the northwestern Doon Valley are
confirmed by geologic mapping as an active fault system that cuts and deforms late
Holocene sediments.

2. The Trans-Yamuna active fault system is subparallel to and cuts the Main Boundary

thrust. The geomorphic south-side-up expression of this fault system indicates that it
does not reactivate the Main Boundary thrust. Instead, it is probably a south-dipping,
high-angle reverse fault with a possible component of strike-slip that in part follows

older thrust zones of weakness.

3. A Paleoseismic trench investigation at Sirmurital did not expose a fault within lower
terrace sediments, but instead revealed the south limb of a late Quaternary syncline
that is parallel to the Sirmurital fault of the Trans-Yamuna active fault system.

4. The presence of an active fault system north of the Indian Himalayan Front thrust
indicates that elastic strain release within the hangingwall is not limited to the front,

but is distributed over a broad area above the decollement. The Trans-Yamuna active
fault system may be one of several secondary hangingwall structures initiated to

accommodate rupture during large magnitude Himalayan earthquakes. Therefore, the


Trans-Yamuna active fault system may contribute toward reducing the slip deficit

between high geodetic and lower geologic convergence rates that are calculated at the
Himalayan front.

51

5.

Strike-slip motion on the Trans-Yamuna active fault system may be related to arcparallel lateral translation of the Karakoram fault block and east-west extension of the
southern Tibet block, in response to oblique convergence between the Indian and
Eurasian plates.

6. A more extensive search for active faults within the Himalaya is necessary to
document the history of earthquakes and constrain recurrence intervals.
Paleoseismology is an important approach toward the mitigation of seismic hazards
associated with large-magnitude earthquakes that rupture the Himalayan front.

52

REFERENCES CITED

Ansari, A.R., Chugh, R.S., Sinvhal, H., Khattri, K.N., and Gaur, V.K., 1976, Geodetic
determination of earth strain and creep on the Krol Thrust in the Dakpathar area,
Dehra Dun District, U.P.: Himalayan Geology, v. 6, p. 323-337.
Arur, M.G., 1982, Results of crustal movement studies in India: Proceedings of the 4th
International Congress, IAEG, v. 8, New Delhi, India.
Arya, A.S., 1996, Vulnerability reduction: key to earthquake protection: Himalayan
Geology, v. 17, p. 33-43.
Auden, J.B., 1934, Geology of the Krol Belt: Records of the Geologic Society of India, v.
67, n. 4, p. 357-454.
Auden, J.B., 1948, Some new limestone and dolomite occurrences in northern India:
Indian Miner, v. 2, n. 2.

Baker, D.M., Lillie, R.J., Yeats, R.S., Johnson, G.D., Yousuf, M., and Zamin, A.S.H.,
1988, Development of the Himalayan frontal thrust zone: Salt Range, Pakistan:
Geology, v. 16, p. 3-7.
Baranowski, J., Armbruster, J., Seeber, L., and Molnar, P., 1984, Focal depths and fault
plane solutions of earthquakes and active tectonics of the Himalaya: Journal of
Geophysical Research, v. 89, p. 6918-6928.
Bhargava, 0.N., 1976, Geology of the Krol belt and associated formations: a reappraisal:
Geological Society of India Memoir, v. 106, part I, p. 167-234.
Bilham, R., 1995, Location and magnitude of the 1833 Nepal earthquake and its relation
to the rupture zones of the contiguous great Himalayan earthquakes: Current
Science, v. 69, n. 2, p. 101-128.

Bilham, R., Blume, F., Bendick, R, and Gaur, V.K., 1998, Geodetic constraints on the
translation and deformation of India: implications for future great Himalayan
earthquakes: Current Science, v. 74, n. 3, p. 213-228.
Bilham, R., Larson, K., Freymueller, J., and Project Idylhim members, 1997, GPS
measurements of present-day convergence across the Nepal Himalaya: Nature, v.
386, n. 6620, p. 61-61.
Boyer, S.E., and Elliot, D., 1982, Thrust systems: American Association of Petroleum
Geologists Bulletin, v. 66, p. 1196-1230.

53

Brookfield, M.E., 1993, The Himalayan passive margin from Precambrian to Cretaceous
times: Sedimentary Geology, v. 84, p. 1-35.

Burbank, D.W., and Raynolds, R.G.H., 1988, Stratigraphic keys to the timing of
thrusting in terrestrial foreland basins: applications to the northwestern Himalaya,
in Kleinspehn, K.L., and Paola, C., eds., New perspectives in basin analysis: New
York, Springer-Verlag, p. 331-351.
Freymueller, J., Bilham, R., Btirgmann, R., Larson, K.M., Paul, J., Jade, S., and Gaur, V.,
1996, Global positioning system measurements of Indian plate motion and
convergence across the Lesser Himalaya: Geophysical Research Letters, v. 23, n.
22, p. 3107-3110.

Gahalaut, V.K., and Chander, R., 1992, On the active tectonics of Dehra Dun region from
observations of ground elevation changes: Journal of the Geological Society of
India, v. 39, p. 61-68.
Gansser, A., 1964, Geology of the Himalayas: New York, Wiley Interscience, 289p.

Hirn, A., Lepine, J.C., Jobert, G., Sapin, M., Wittlinger, G., Xin, X.Z., Yaun, G.E., Jing,
W.X., Wen, T.J., Bai, X.S., Pandey, M.R., and Tater, J.M., 1984, Crustal
structure and variability of the Himalayan border of Tibet: Nature, v. 307, p.23-25.
Hodges, K.V., Parrish, R.R., Housh, T.B., Lux, D.R., Burchfiel, B.C., Royden, L.H., and
Chen, Z., 1992, Simultaneous Miocene extension and shortening in the Himalayan
orogen: Science, v. 258, p. 1466-1470.
Jackson, M., and Bilham, R., 1994, Constraints on Himalayan deformation inferred from
vertical velocity fields in Nepal and Tibet: Journal of Geophysical Research, v. 99,
p. 13, 897- 13,912.
Kanamori, H., 1977, The energy release in great earthquakes: Journal of Geophysical
Research, v. 82, p. 2981-2987.

Kapur, R.B.S.K.C., 1934, Gazetteer of the Sirmur State Part A: Punjab, India, Punjab
Government, 129p.
Karunakaran, C., and Ranga Rao, A., 1976, Status of exploration for hydrocarbons in the
Himalayan regioncontributions to stratigraphy and structure: Himalayan
Geology, Section III, New Delhi: Geological Survey of India, Miscellaneous
Publication, v. 1, n. 5, p. 1-66.
Kelleher, J.A, Sykes, L.R., and Oliver, J., 1973, Possible criteria for predicting earthquake
locations and their applications to major plate boundaries of the Pacific and
Caribbean: Journal of Geophysical Research, v. 78, p. 2547-2585.

54

Khan A.A., and Dubey U.S., 1981, Evolution of Quaternary landscape of intermontane
Yamuna Valley in Dehra Dun (U.P.) and Sirmur (H.P.): Records of the Geological
Survey of India, v. 112, p. 125-135.

Khattri, K.N., 1987, Great earthquakes, seismicity gaps and potential for earthquake
disaster along the Himalayan plate boundary: Tectonophysics, v. 138, p. 79-92.

Khattri, K.N., Chander, R., Gaur, V.K., Sarkar, I., and Kumar, S., 1989, New
seismological results on the tectonics of the Garwhal Himalaya: Proceedings of the
Indian Academy of Sciences (Earth and Planetary Sciences), v. 98, n. 1, p. 91-109.
Khattri, K.N., and Tyagi, A.K., 1983, Seismicity patterns in the Himalayan plate boundary
and identification of areas of high seismic potential: Tectonophysics, v. 96, p. 281297.

Krishnaswamy, V.S., Jalote, S.P., and Shome, S.K., 1970, Recent crustal movements in
northwest Himalaya and the Gangetic foredeep and related patterns of seismicity:
Proceedings of the 4th Symposium of Earthquake Engineering, Roorkee.
Lave, J., and Avouac, J.P., 1998a, Abandoned fluvial terraces across the Siwalik Hills
(Nepal): I, Fluvial Response to Climatic and Tectonic Forcing: Journal of
Geophysical Research, (in press).
Lave, J., and Avouac, J.P., 1998b, Abandoned fluvial terraces across the Siwalik Hills
(Nepal): II, Active fault-bend-folding at the MFT and implications for Himalayan
seismotectonics: Journal of Geophysical Research, (in press).
Le Fort, P., 1975, Himalayas: the collided range. Present knowledge of the continental
arc: American Journal of Science, v. 275-A, p. 1-44.

Leathers, M., 1987, Balanced structural cross-section of the western Salt range and
Potwar Plateau: deformation near the strike-slip terminus of an overthrust sheet:
Master of Science thesis, Oregon State University, Corvallis, Oregon, 271 p.
Lyon-Caen, H., and Molnar, P., 1985, Gravity anomalies, flexure of the Indian Plate, and
the structure, support and evolution of the Himalaya and Ganga Basin: Tectonics,
v. 4, n. 6, p. 513-538.
McCaffrey, R., 1992, Oblique plate convergence, slip vectors, and forearc deformation:
Journal of Geophysical Research, v. 97, p. 8905-8915.
McCaffrey, R., and Nabelek, J, 1998, Role of oblique convergence in the active
deformation of the Himalayas and southern Tibet plateau: Geology, v. 26, n. 8, p.
691-694.

55

McCalpin, J.P., and Nelson, A.R., 1996, Introduction to Paleoseismology, in McCalpin,


J.P., ed., Paleoseismology: San Diego, Academic Press, p. 1-28.

Meigs, A. J., Burbank, D. W., and Beck, R. A., 1995, Middle-late Miocene (ca. 10 Ma)
formation of the Main Boundary thrust in the western Himalaya: Geology, v. 23, p.
423-426.
Meijerink, A.M.J., 1974, Doon Valley, a photo-hydrological reconnaissance survey:
Enschede Special Publication of the International Training Center, the
Netherlands:, p. 283-308.
Miall, A.D., 1996, The geology of fluvial deposits: sedimentary facies, basin analysis, and
petroleum geology: Berlin, Springer- Verlag, 582p.

Middlemiss, C.S., 1905, Preliminary account of the Kangra earthquake of 4th April, 1905:
Records of the Geologic Survey of India, v. 33, p. 258-294.
Middlemiss, C.S., 1910, Kangra earthquake of 4 April 1905: Memoirs of the Geological
Survey of India, v. 38, 409p.

Mohindra, R., and Thakur, V.C., 1998, Historic large earthquake-induced soft sediment
deformation features in the Sub-Himalayan Doon Valley: Geological Magazine, v.
135, n. 2, p. 269-281.
Molnar, P., 1987, Inversion of profiles of uplift rates for the geometry of dip-slip faults at
depth, with examples from the Alps and Himalaya: Armales Geophysicae, v. 5B, n.
6, p. 663-670.
Molnar, P., 1990, A review of seismicity and the rates of active underthrusting and
deformation at the Himalaya: Journal of Himalayan Geology, v. 1, n. 2, p. 131154.

Molnar, P., and Tapponnier, P., 1975, Cenozoic tectonics of Asia: effects of a continental
collision: Science, v. 189, p,419-426.
Najman, Y.M.R., Enkin, R.J., Johnson, M.R.W., Robertson, A.H.F., and Baker, J., 1994,
Paleomagnetic dating of earliest continental Himalayan foredeep sediments:
implications for Himalayan evolution: Earth and Planetary Science Letters, v. 128,
p. 713-718.
Najman, Y.M.R., Pringle, M.S., Johnson, M.R.W., Robertson, A.H.F., and Wijbrans, J.R.,
1997, Laser 40Ar/39Ar dating of single detrital muscovite grains from early foreland basin sedimentary deposits in India: implications for early Himalayan
evolution: Geology, v. 25, n. 6, p. 535-538.

56

Nakata, T., 1972, Geomorphic history and crustal movements of the foothills of the
Himalayas: Science Reports of the Tohoku University, 7th Series (Geography), v.
22, n. 1, 177p.
Nakata, T., 1982, A photogrammetric study on active faults in the Nepal Himalayas:
Journal of Nepal Geologic Society, v. 2, p. 67-80.
Nakata, T., 1986, Active faults of the Nepal Himalayas and their tectonic significance:
Proceedings of the International Symposium on Neotectonics in South Asia, Dehra
Dun, India, February 18-21, p. 127-136.
Nakata, T., 1989, Active faults of the Himalaya of India and Nepal: Geological Society of
America Special Paper 232, p. 243-264.

Nakata, T., Iwata, S., Yamanaka, H., Yagi, H, and Maemoku, H., 1984, Tectonic
landforms of several active faults in the western Nepal Himalayas: Journal of Nepal
Geologic Society, Special Issue, v. 4, p. 177-200.
Ni, J., and Barazangi, M., 1984, Seismotectonics of the Himalayan collision zone:
geometry of the underthrusting Indian plate beneath the Himalaya: Journal of
Geophysical Research, v. 89, n. B2, p. 1147-1163.

Nossin, J.J., 1971, Outline of the geomorphology of the Doon Valley, northern U.P.,
India: Zeitschrift far Geomorphologie, v. 12, p. 18-50.
Oldham, T., 1882, A catalogue of Indian earthquakes from the earliest time to the end of
A.D. 1869: Memiors of the Geological Survey of India, v. 19, p. 163-215.

Parkash, B., Sharma, R.P., and Roy, A.K., 1980, The Siwalik Group (molasse)Sediments
shed by collision of continental plates: Sedimentary Geology, v. 25, p. 127-159.
Patriat, P., and Achache, J., 1984, India-Eurasia collision chronology has implications for
crustal shortening and driving mechanism of plates: Nature, v. 311, p. 615-621.
Peltzer, G., and Saucier, F., 1996, Present day kinematics of Asia derived from geologic
fault rates: Journal of Geophysical Research, v. 101, n. B12, p. 27,943-27,956.

Pennock, E.S., Lillie, R.J., Zamin, A.S.H., and Yousuf, M., 1989, Structural interpretation
of seismic reflection data from eastern Salt Range and Potwar Plateau: American
Association of Petroleum Geologists Bulletin, v. 73, p. 841-857.
Philip, G., 1996, Landsat Thematic Mapper data analysis for Quaternary tectonics in parts
of the Doon Valley, NW Himalaya, India: International Journal of Remote
Sensing, v. 17, n. 1, p. 143-153.

57

Powers, P.M., Lillie, R.J., and Yeats, R.S., 1998, Structure and shortening of the Kangra
and Debra Dun reentrants, Sub-Himalaya, India: Geological Society of American
Bulletin, v. 110, n. 7, p. 1010-1027.
Quittmeyer, R.C., and Jacob, K.H., 1979, Historical and modern seismicity of Pakistan,
Afghanistan, northwestern India, and southwestern Iran: Bulletin of the
Seismological Society of America, v. 69, n. 3, p. 773-823.
Raiverman, V., Ganju, J.L., and Misra, V.N, 1976, Geological map of the Himalayan
foothills between the Rivers Ravi and Yamuna: Oil and Natural Gas Commission
of India Unpublished Report.
Raiverman, V., Ganju, J.L. and Misra, V.N., 1981, A new look into the stratigraphy of
Cenozoic sedimentation of the Himalayan foothills between the Ravi and Yamuna
rivers: Geological Survey of India, v. 41, p. 233-246.
Raiverman, V., Kunte, S.V., and Mukherjea, A., 1983, Basin geometry, Cenozoic
sedimentation and hydrocarbon prospects in NW Himalaya and Indo-Gangetic
plains: Petroleum Asia Journal, Special Volume 1, p. 67-92.

Rao, D.P., 1977, A note on recent movements and origin of some piedmont deposits of
Debra Dun Valley: Photonirvachak, v. 5, n. 1, p. 35-40.

Rao, D.P., 1978, Geomorphology and neotectonicsan example from Sub-Himalaya:


Proceedings of the symposium on morphology and evolution of landforms,
Geology Department, Delhi University, p. 88-98.
Rao, D.P., 1986, An approach to the quantitative estimates of rates of recent movements:
the Sub-Himalayan example: Royal Society of New Zealand, Bulletin 24, p. 507517.

Rockwell, T. K., Keller, E.A., Clark, M.N., and Johnson, D.L., 1984, Chronology and
rates of faulting of Ventura River terraces, California: Geological Society of
America, v. 95, p. 1466-1474.
Schelling, D., and Arita, K., 1991, Thrust tectonics, crustal shortening, and the structure
the far-eastern Nepal Himalaya: Tectonics, v.10, p. 851-862.
Seeber, L, and Armbruster, J.G., 1981, Great detachment earthquakes along the
Himalayan arc and long-term Forecasting, in Simpson, D.W., and Richards, P.G.,
eds., Earthquake prediction: an international review, Volume Maurice Ewing
Series: 4): Washington, D.C., American Geophysical Union, p. 259-277.

58

Seeber, L., Armbruster, J.G., and Quittmeyer, R.C., 1981, Seismicity and continental
subduction in the Himalayan arc, in Gupta, H.K., and Delaney, F.M., eds., Zagros,
Hindu Kush, Himalaya, geodynamic evolution: American Geophysical Union
Geodynamic Series, v. 3, p. 215-242.

Sinhval, H.S., Agarwal, P.N., King, G.C.P., and Gaur, V.K., 1973, Interpretation of
measured movement at a Himalayan (Nahan) thrust: Geophysical Journal of the
Royal Astronomical Society, v. 34, p. 203-210.
Srikantia, S.V., and Sharma, R.P., 1976, Geology of the Shaft belt and adjoining areas:
Geological Society of India Memoir, v. 106, part I, p. 31-166.

Stuiver, M., and Reimer, P.J., 1993, Extended "C data base and revised CALIB 3.0 14C
age calibration program: Radiocarbon, v. 35, p.215-230.
Swamy, M.M., 1986, Neotectonic movements across thrust planes and younger fault
planes in Gin Hydroelectric project area, District Sirmur, Himachal Pradesh, India:
Proceedings (Post Session) International Symposium on Neotectonics in south
Asia, Dehra Dun, India, Feb. 18-21., p. 58-67.
Sykes, L.R., 1971, Aftershock zones of great earthquakes, seismicity gaps, and
earthquake prediction for Alaska and the Aleutians: Journal of Geophysical
Research, v. 76, p. 8021-8041.
Tapponnier, P., and Molnar, P., 1977, Active faulting and tectonics in China: Journal of
Geophysical Research, v. 82, n. 20, p 2905-2930.
Thakur, V.C., 1992, Geology of the Western Himalaya: New York, Pergamon Press,
366p.

Thomas, R.G., Smith, D.G., Wood, J.M., Visser, J., Calvarerly-Range, E.A., Koster,
E.H., 1987, Inclined heterolithic stratificationterminology, description,
interpretation and significance: Sedimentary Geology, v. 53, p. 123-179.
Valdiya, K.S., 1976, Himalayan transverse faults and folds and their parallelism with
subsurface structures of north Indian plains: Tectonophysics, v. 32, p. 353-386.
Valdiya, K.S., 1980, The intracrustal boundary thrusts of the Himalaya:
Tectonophysics, v. 63, p. 323-348.

Valdiya, K.S., 1986, Neotectonic movements in the Himalayan belt: Proceedings of the
International Symposium on Neotectonics in South Asia, Dehra Dun, India, Feb.
18-21, p. 241-251.

59

Valdiya, K.S., 1992, The Main Boundary Thrust Zone of the Himalaya, India: Annales
Tectonicae Special Issue, Supplement to v.6, p.54-84.

Vohra, C.P., Raina, A.K., Dua, K.J.S., and Khanna, P.C., 1976, The outer Pre-Tertiary
limestone belt between Tundapathar and Sataun: Geological Society of India
Memoir, v. 106, part I, p. 3-30.
Wells D.L., and Coppersmith, K.J., 1994, New empirical relationships among magnitude,
rupture length, rupture area, and surface displacement: Seismological Society of
America Bulletin, v. 84, p. 974-1002.

Wesnousky, S. G., Kumar, S., Mohindra, R., and Thakur, V.C., 1998, Holocene slip rate
of the Himalayan Frontal thrust of India: unpublished preprint.
Yeats, R.S., 1987, Late Cenozoic structure of the Santa Susana fault zone; United States
Geological Survey Professional Paper 1339, p. 137-160.
Yeats, R.S., and Hussain, A., 1987, Timing of structural events in the Himalayan foothills
of northwest Pakistan: Geological Society of America Bulletin, v. 99, p. 161-176.
Yeats, R.S., and Lillie, R.J., 1991, Contemporary tectonics of the Himalayan Frontal fault
system: folds, blind thrusts, and the 1905 Kangra earthquake: Journal of Structural
Geology, v. 13, n. 2, p. 215-225.

Yeats, R.S., Nakata, T., Farah, A., Fort, A., Mirza, M.A., Pandey, M.R., and Stein, R.S.,
1992, The Himalayan frontal fault system: Annales Tectonicae, Special Issue, v.6,
p.85-98.
Yeats, R.S., and Thakur, V.C., 1998, Reassessment of earthquake hazard based on a faultbend fold model of the Himalayan plate-boundary fault: Current Science, v. 74, n.
3, p. 230-233.
Zhao, W., Nelson, K.D., and project INDEPTH team, 1993, Deep seismic reflection
evidence for continental underthrusting beneath southern Tibet: Nature, v.366, p.
557-559.

Вам также может понравиться