Sei sulla pagina 1di 130

AN ABSTRACT OF THE THESIS OF

Dan M. Baker for the degree of Master of Science in Geology presented


on June 4, 1987.
Title: Balanced Structural Cross-Sections of the Central Salt Range and
Potwar Plateau of Pakistan: Shortening and Overthrust Deformation

Abstract approved:
Redacted for Privacy
5obert J. Lillie

The Salt Range and Potwar Plateau are part of the currently active
foreland fold-thrust belt of the Himalaya in Pakistan. Seismic reflec-
tion data have been combined with surface mapping, drillhole and
paleomagnetic information to construct balanced structural cross-sec-
tions of the current frontal thrust zone (Salt Range), and to estimate
the amount of horizontal contraction and timing of deformation for the
central Salt Range/Potwar Plateau.
The data show that a 1 km offset of the basement acted as a
buttress that caused the central Salt Range/Potwar Plateau thrust sheet
to ramp to the surface, exposing Mesozoic and Paleozoic strata at the
range front. The basement offset may have been formed during Eocam-
brian rifting; however, an origin related to Neogene crustal flexure is
supported. The frontal portion of the thrust sheet was passively
folded as it overrode the subthrust surface upon a thick, ductile layer
of salt. The proposed ramping mechanism includes deformation prior to
overriding the basement offset and suggests a smoother thrust trajec-
tory than the ideal, stairstep profile. Balanced sections do not
indicate a lateral change in shortening in the central Salt Range, as
would be expected during proposed regional rotation during ramping.
Just north of the Salt Range, the thrust sheet is undeformed,
except for minor folding related to salt redistibution. Lack of
internal deformation in the Soan syncline is due to decoupling of
sediments from the basement along the thick Eocambrian salt layer.
Still farther north, in the Northern Potwar Deformed Zone (NPDZ), a
hinterland dipping duplex is disharmonic with mapped surface structure,
suggesting an upper-level detachment. Duplexing ends southward in a
"triangle zone" geometry, and an apparent imbricate fan. Beneath the
northern limb of the Soan syncline, the duplex and imbricate fan are
separated by an oblique ramp. Section balancing suggests about 45 km
of shortening across the NPDZ.
The timing of structural events is constrained by stratigraphic
evidence, paleomagnetic dating of unconformities and sediment accumula-
tion rates. Early to middle Pliocene (approx. 4.5 Ma) clastic deposi-
tion in the southern Potwar Plateau, previously interpreted in terms of
compressional deformation, may instead be due to passage of the peri-
pheral bulge (Sargodha High) through the Indian lithospheric plate
causing basement normal faulting, uplift, and erosion. Although the
timing of deformation in the central Potwar Plateau is poorly con-
strained, the Salt Range thrust sheet is estimated to have overridden
the basement offset from 2.1-1.6 mya. Cross-section balancing gives a
minimum of 19 to 23 km of shortening across this frontal ramp. The
rate of Himalayan convergence which can be attributed to underthrusting
of Indian basement beneath sediments in the Pakistan foreland is there-
fore 9-14 mm/yr. Deformation front migration appears to be controlled
by the distribution of salt and basement buttressing in the central
Salt Range/Potwar Plateau. Rapid facies progradation, measured in the
Jhelum Reentrant, is thought to be controlled by thrust propagation
rates and not indicative of typical foredeep migration. Restorations
suggest that the lower and middle Siwalik molasse depocenter was in the
southern Potwar Plateau, prior to jumping south of the Salt Range about
2 mya.
Balanced Structural Cross-Sections

of the Central Salt Range and Potwar Plateau of Pakistan

Shortening and Overthrust Deformation

by

Dan M. Baker

A THESIS

submitted to

Oregon State University

in partial fulfillment

of the requirements for the

degree of

Master of Science

Completed June 4, 1987

Commencement June, 1988


APPROVED:

Redacted for Privacy


Professor of Geoeogy in charge of major

Redacted for Privacy


Cailan of Depar ent of Geology

Redacted for Privacy

Dean of Gradua School

Date thesis is presented June 4, 1987


ACKNOWLEDGEMENTS

Thesis work in northern Pakistan was supported by NSF grants INT-


8118403, EAR-8318194, INT-8609914, and EAR-8608224. Additional assis-
tance was provided by Amoco Production Company and Chevron Inc. I am
grateful to the Oil and Gas Development Corporation and the Ministry
of Petroleum and Natural Resources of Pakistan for the use of reflec-
tion and gravity data discussed in this study. Depth conversion of
seismic sections performed with the AIMS seismic modeling package,
donated to Oregon State University by Geoquest, International.
Thanks to Bob Lillie for his time, help in writing, and enthusi-
asm. The time and consideration given by other committee members,
R. Yeats, V. Kulm and D. Bibee, plus M. Yousuf was appreciated. I

would also like to thank my colleagues in this study for their help and
willingness to share ideas, especially M. Leathers whose hard work got
things going.
I am grateful to Mrs. Hamerick, a grammar school teacher, whose
special attention got me interested in science. The moral support of
my parents and family members helped me see past the poverty that
accompanied graduate school. Special thanks to my wife, Linda,
for her assistance, especially in typing, motivation to return to
school, and her support throughout the process.
TABLE OF CONTENTS

Page
INTRODUCTION 1

Objectives 1

Previous Work 2

Tectonic Setting 6
Methods 9
Depth Sections 9
Balanced Sections 10
Restorations 16
Calculation of Shortening 17

STRATIGRAPHY AND SEDIMENTATION 19


Stratigraphy 19
Sedimentation and Depositional Models 21
Depotectonics 22
Depocenter Migration across the Central SR/PP 22
Foredeep Migration Rates 23

STRUCTURE OF THE CENTRAL SALT RANGE AND POTWAR PLATEAU 29


General Description 29
Northern Potwar Deformed Zone 32
Description 32
Analysis 41
Transition from the NPDZ to the Soan Syncline 44
Description 44
Analysis 45
Southern Potwar Plateau/Soan Trough 45
Transition from the Southern Potwar Plateau
to the Salt Range 49
Description 49
Analysis 49
Salt Range 58
General Description 58
Eastward Termination of the Salt Range Thrust Sheet 61
Structures within the Central Salt Range 64
Ramping of the Salt Range Thrust Sheet 66
Mechanism 66
Removal of Molasse from Beneath the Salt Range 68

TIMING CONSTRAINTS ON THE EVOLUTION OF THE FOLD-THRUST BELT 72


Migration of Folding and Thrusting into the Foreland 72
Decollement Propagation Rate Calculation 73
Motion of the Thrust System 74
Tectonic Rotations 75
Timing of Basement Faulting 76

SHORTENING AND ITS IMPLICATIONS 82


Amount of Shortening 82
Salt Range 82
Southern Potwar Plateau/Soan Trough 82
Northern Potwar Plateau 83
Frontal Shortening Rate and Comparison to Plate
Convergence Rates 83
Tectonic Rotation Versus the Distribution of Shortening 85

CONCLUSIONS 87
Future Work 89

BIBLIOGRAPHY 90

APPENDICES 101
(A) Development of the Himalayan Frontal Thrust Zone:
Salt Range, Pakistan 101
(B) Restoration Results and Implications 116
(C) Factors Influencing Shortening Calculations 118
(D) Petroleum Geology 120
LIST OF FIGURES

Figure Page

1. Regional location map. 3

2. Tectonic map of northern Pakistan. 5

3. Map of seismic and well control. 8

4. Example of velocities used for seismic time to


depth conversion. 11

5. Structure map of the SR/PP with paleomagnetic


sample locations. 14

6. Sedimentation model, showing southward shift of


molasse depocenter. 24

7. Restored positions of the Sargodha high, Salt


Range, basement offset and Baun locality
at 4.5 mya. 26

8. Block diagram illustrating a change in


deformation sytle across an oblique ramp
in the northern Potwar deformed zone. 30

9. Seismic transect across the northern Potwar


deformed zone west of the oblique ramp. 33

10. Seismic transect across the northern Potwar


deformed zone east of the oblique ramp. 35

11. Seismic transect across the northern limb of


the Soan syncline. 37

12. Summary of deformation sytles in the northern


Potwar Plateau. 40

13. Regional balanced cross-section and


restorations. 42

14. Seismic transect across the southern


Potwar Plateau. 46

15. Seismic profile across the frontal basement


ramp. 50

16. Geologic map of the central Salt Range area. 51

17. Frontal balanced cross-sections. 52


18. Frontal ramping diagram showing preramping
deformation, ramp shallowing, and footwall
collapse. 56

19. Block diagram illustrating evolution of the


central Salt Range. 59

20. Hangingwall/footwall cutoff separation map


of the east-central Salt Range. 62

21. Restorations illustrating stratigraphic thickness


changes versus the age of basement faulting. 69

22. Schematic diagram illustrating allochthonous and


autochthonous origin of Upper Siwalik
conglomerates preserved at the Baun locality. 79

23. Schematic crustal section demonstrating potential


distribution of convergence absorbtion. 84

Regional location map. 103


Al.

Structure/geology map of the SR/PP. 105


A2.

A3. Schematic cross-section of the Himalayan foreland


fold-thrust belt in Pakistan. 107

Seismic profile of the frontal ramp region. 109


A4.

A5. Balanced cross-section and restorations demon-


strating a possible mode of origin of the
Salt Range. 112
BALANCED STRUCTURAL CROSS-SECTIONS
OF THE CENTRAL SALT RANGE AND POTWAR PLATEAU OF PAKISTAN:
SHORTENING AND OVERTHRUST DEFORMATION

INTRODUCTION

The active western Himalayan foreland fold-thrust belt presents an


opportunity to study a developing stage of collision. Approximately
3000 km of seismic reflection profiles (supplied to Oregon State
University by the Pakistan Oil and Gas Development Corporation and the
Ministry of Petroleum), have been interpreted in conjunction with
surface maps (especially those of Gee, 1980), and published and unpub-
lished well and gravity data. These data were used to construct
balanced structural cross-sections of the central Salt Range and Potwar
Plateau and to estimate amounts of shortening. The same data set was

used by Leathers (1987) and Pennock (in prep) to construct cross-sec-


tions to the west and east of the thesis area, by Jaume (1986) to study
the mechanics of thrusting and by Duroy (1986) to analyse lithospheric
flexure for the Himalaya of Pakistan. Appendix A (Baker et al., in
review) is a brief overview of portions of this study, concentrating on
the timing of frontal ramping and the calculation of a frontal shorten-
ing rate.

Objectives

The Pakistani foreland contains a wide variety of structures


related to the combined effects of compressional and salt tectonics. A

major objective of this study is to describe structural styles and


examine the relationships between different structures. In addition,

recent sedimentological and stratigraphic studies in the Salt Range/


Potwar Plateau (SR/PP) area have utilized magnetostratigraphy and
tephrachronology to age-calibrate individual horizons in the Siwalik
molasse (e.g Johnson et al., 1986). These data constrain the timing of
individual deformational events and the progression of the thrust front
across the foreland basin. When used in combination with structural
information, the evolution of the fold-thrust belt, especially the Salt
2

Range, can be inferred. The effect of the progressing deformation


front and basement faulting on foredeep sedimentation is also consid-
ered. An attempt is made to solve the following problems: (1) the
origin of the down-to-the north, basement offset that controls Salt
Range ramping; (2) how ramping in the Salt Range compares with previ-
ously developed models of thrust ramping; (3) the role of thrust sheet
ramping in the overall rotation documented in the area; (4) the amount
of horizontal contraction of cover rocks; and (5) the rate of short-
ening in the frontal portion of the Pakistani foreland fold-thrust
belt.

Previous Work

The Salt Range, with its excellent exposures of a wide range of


Phanerozoic strata, has long attracted the interests of geologists. In

addition to the stratigraphic significance of its outcrops, the pre-


sence of highly deformed Mesozoic and Paleozoic rocks so far outboard
of the major Himalayan deformation zone has attracted the attention of
structural geologists.
Stratigraphic classification of the Tertiary Siwalik succession
across the SR/PP is an ongoing problem. Maps and general accounts of
the stratigraphy were prepared by Wadia (1928), Cotter (1933), and Gill
(1952). More recently, attempts at an absolute chronostratigraphy
through the application of paleomagnetic stratigraphy and fission track
dating, have been made (e.g. Barndt et al., 1978).
Detailed maps (1:50,000), recently published by Gee (1980), demon-
strate the structural complexity of the Salt Range. Attempts to
unravel the history of less spectacular structures in the Potwar
Plateau were made by Pinfold (1918), Cotter (1933), Gill (1952), Martin
(1962) and later by Voskrensenskiy (1978). A summary of some of the
data and results from subsurface studies undertaken during hydrocarbon
exploration can be found in Khan et al. (1986).
The significance of the position of the SR/PP on the west side of
the bend in the Himalayan trend at the Hazara-Kashmir Syntaxis (HKS;
fig. 1), and on the east end of the sinuous Pakistani fold belt remains
a problem. The Ceozoic festoons in both India and Pakistan are thought
Fig. 1: Regional location map, showing position of the Salt Range/Pot-
war Plateau (SR/PP) near the west end of the overall Himalayan
trend. Sinuous lines through southern and central Pakistan
represent the Pakistan fold-thrust belt. HKS=Hazara-Kashmir
Syntaxis; ISZ=Indus Suture Zone; MBT=Main Boundary Thrust;
MKT= Main Karakorum Thrust; SH=Sargodha High. Rectangular box
denotes area of figure 2.
4

to be controlled by the reactivation of Precambrian basement lineaments


by Swaminath and others (1964), or by the distribution of evaporites at
the base of the sedimentary sequence (Seeber et al., 1981). A more
general account concerning possible causes of the scalloped outline of
Pakistan's mountain fronts was given by Sarwar and DeJong (1979).
The Hazara region (including the SR/PP) was genetically shown to
be related to the Himalaya by Wadia (1931), who noted that the rocks of
the Hazara are a stratigraphic continuation of those exposed in the
Himalaya. He suggested that the two belts are sections of the same
mountain system, bent around a horst-like projection of the Indian
shield. Crawford (1974) offered an alternative, hypothesizing that the
Salt Range has been rotated 75° counterclockwise from a position in
line with the main Himalayan front. His theory is in part supported by
recent work which shows that thrusting in the foreland is accompanied
by counterclockwise rotation of portions of the Salt Range (Opdyke et
al., 1982). A conflicting result, however, was presented by Powell
(1979), who showed that the Salt Range and the main Himalaya have an
east-west component of convergence, suggesting that the HKS may be the
collision point of two different parts of the orogen. Davis and
Engelder (1985) suggest that relative movements between the Salt Range
and the Himalaya can also be explained by the lubricating effects of
salt within the decollement zone west of the syntaxis.
Recent work by Yeats and others (1984), who documented that
deformation in the eastern Salt Range is younger than 0.4 mya, and by
Lillie and Yousuf (1986), who explained the emplacement of the Salt
Range by ramping over a basement fault, have set up this investigation
into the evolution of the central Salt Range. A gross balanced section
across this area (Coward and Butler, 1985) provides a basic framework
for investigation, but has the main decollement at the base of the
molasse instead of within the Eocambrian evaporites, as shown on
reflection profiles and includes several thrusts not rocognized in the
field (Yeats and Hussain, in press). Previous estimates of shortening
represented by overthrusting at the Salt Range are 35 km by Gee (1934)
and 50 km by Sarwar and DeJong (1979). An estimate of total shortening
from the Salt Range to the Main Mantle Thrust (fig. 2) is 470 km
(Coward and Butler, 1985).
5
7Q° 74°
37°69' °

NORTHERN PAKISTAN
OC

Generalized
Tectonic Map Thrust
(After Kazmi and Rana, 1982)
LEGEND (FOR AREA SOUTH OF M B T
36° Neogene Malone

Chitral.
Paleocene to Eocene
Monne to Continental
Penman to Jurassic
Sedimentary Rocks KOHISTAN
.1111PrIcantbrian M Campion Sedimentary Room ARC
12c1 Preeartaxian Indian Shook'

Oregon State University


1986
35°

Fire Line

Ka*/ Rivav
KoDai
AJottod PESHAWAR
J!,!'
or abad sr=go.2.
34°
Pesnowor

B. 1.
ono,

POTWAR
Soon Rnn

33° PLATEAU

32°

PLAIN
Lyallpur

3'°
Fig. 2: Generalized tectonic map of northern Pakistan, illustrating
features in, and north of the western Himalayan foreland (after
Kazmi and Rana, 1982). Salt Range is the area of Precambrian to
Eocene sedimentary rocks north of the Salt Range Thrust. Dotted
line north of the Salt Range marks the subsurface position of the
basement offset. Bold rectangles represent the areas of figures
showing available seismic coverage. Right rectangle denotes area
of figure 3. Left rectangle represents western SR/PP coverage
(see Leathers, 1987). X-X' and Y-Y' are cross-sections of Leath-
ers (1987) and Z-Z' is a crustal cross-section (see Duroy, 1986;
Duroy et al., in press). HR=Hill Ranges; JR=Jhelum Reentrant;
KH=Kirana Hills; MR=Mianwali Reentrant; and SuR=Surghar Range.
6

Tectonic Setting

The Himalaya of Pakistan is a product of the ongoing collision of


the Indian subcontinent with Eurasia. The continent-continent colli-
sional event began in Late Eocene, after docking of the Kohistan/Ladakh
arc (fig. 2) with Asia in late Cretaceous (Windley, 1983). Seafloor
reconstructions indicate that since the collision began, 1500 to 3500
km of convergence between Eurasia and India has occurred. Convergence
is accommodated by giant underthrusting of one continental block
beneath the other and/or diffuse deformation over a broad zone (Molnar
and Tapponnier, 1975; Tapponnier et al., 1982). Some underthrusting of
the Indian subcontinent beneath Eurasia is supported by reconstructions
of Gondwana, and by gravity and seismic refraction studies indicating
twice the normal thickness of continental crust beneath the high Hima-
laya in Tibet and northern Pakistan (Finetti et al., 1981). Diffuse
deformation involves crustal thickening via folding and thrusting, and
lateral strain by strike-slip faulting (Molnar and Tapponnier, 1975).
Continental fragments north of the Himalaya may be squeezed out of the
way of the advancing Indian subcontinent along strike-slip faults
("escape block tectonics").
To the south, in the lesser Himalaya of northern Pakistan (Hill
Ranges; fig. 2), detachments at upper crustal levels occur along a
series of south-verging thrusts (Yeats and Lawrence, 1984). The

southern margin of the collision zone in Pakistan (Salt Range, Kohat


and Potwar Plateaus) is an active foreland fold-thrust belt formed in
response to the underthrusting of cratonic India beneath its own

Phanerozoic sedimentary cover. On the basis of microseismicity, the


Salt Range, Potwar Plateau and Hill Ranges are interpreted as thin
skinned, underlain by a low-angle decollement which surfaces at the
Salt Range front (Seeber and Armbruster, 1979).
At the Hazara-Kashmir Syntaxis, the fold-thrust belt changes in
trend and broadens in Pakistan (fig. 1). In India east of the Punjab
reentrant, there is strong coupling between sediments and basement;
consequently the zone of underthrusting is narrow and has a high angle
of cross-sectional taper (Davis et al., 1983; Davis and Engelder,
1985). In Pakistan, on the west flank of the Hazara-Kashmir Syntaxis,
7

low-strength evaporites of late Precambrian to Early Cambrian age


constitute the zone of decollement (Crawford, 1974). As a result and
in contrast to India, the zone of underthrusting extends far out over
the foreland (Seeber and Armbruster, 1979), more than 100 km south of
the Main Boundary Thrust (fig. 2), and it has a narrow cross-sectional
taper. Further evidence for the decoupling properties of the evapor-
ites lies in the large detachment earthquakes which occur beneath the
foreland in India, but not in Pakistan (Quittmeyer and Jacob, 1979;
Seeber et al., 1981).
In northern Pakistan, a southward progression of major compres-
sional deformation has been documented (Raynolds and Johnson, 1985;
Johnson et al., 1986), although out of sequence deformation also occurs
(Burbank and Raynolds, in press). At the southern margin of the
Peshawar basin, major thrusting occurred prior to 2.8 mya (Burbank,
1982), while in the eastern Salt Range, major deformation may be
younger than 0.4 mya (fig. 2; Johnson, et al., 1979). The deformation
front in the central Salt Range may now be shifting to the Lilla
structure, a thrust-cored anticline in the Jhelum River plain 15 km
south of the Salt Range.
South of the fold-thrust belt, sediments slope gently northward
from the Sargodha High, an area of moderate seismicity, including

exposures of Precambrian rocks at the Kirana Hills (fig. 2; Farah et


al., 1977). It is often assumed that the Sargodha High is an old
Precambrian feature analogous to the Aravalli Range of India. However,

similar stratigraphy in wells drilled north and south of the Sargodha


High, plus sharp topographic breaks between the Kirana hills and the
surrounding plains, suggest that this ridge is a relatively young
feature (Yeats and Lawrence, 1984). This ridge parallels the overall
northwest-southeast Himalayan trend and is perhaps an "outer swell"
related to the downflexing of the underthrusting Indian plate (Seeber
et al., 1981; Duroy, 1986; Duroy et al., in press). Alternatively, the

Sargodha High could represent an incipient Himalayan basement thrust


analogous to the Main Central Thrust (MCT) in India (Yeats and Law-
rence, 1984).
In addition to the Himalayan orogeny, evidence of past and contin-
uing epeirogenic movements exist in the stratigraphic sections of the
8

+3,5 7 3' 61 745


7.5

33.45.

53. 06

010,A
490010,
33,5

33,X

3245

HIMALAYAN FORELAND THRUST BELT


NORTHERN PAKISTAN

SALT RANGE POTWAR PLATEAU


E xplorofion 119 .1 *619 and 0B A

5666c Lines
.. air 4a. C.00.000 004.000C.CI . San P1.00. row Ca0n
32.15'
O. 4a. 01000.40 Pomo. eso.. %ow Kxxx
9
0 69 060 99666 C. o
Amos. 0100. 0...00. Cs. sre, llo 0080
1006.0.

6600*.L..Nmaavol
SAR000A

Down Mar um.soy 1966


31500'
73.15' 73.30' 73.45' '4 00
7205 7e 30' r295 731;,0

Fig. 3: Map of available seismic and well control in the central and
eastern SR/PP. See Khan et al. (1986) for additional information
and Leathers (1987) for well data and structure contour maps.
Portions of seismic lines illustrated in text (figs. 9, 10, 11,
14, and 15) and used to construct balanced cross-sections are
figure 5.
denoted by solid lines. Rectangular box denotes area of
9

Salt Range, Potwar Plateau and Jhelum reentrant. The concentration


of Eocambrian evaporites into distinct basins, uplift of the western
Salt Range region relative to the eastern Salt Range during or after
the Cambrian, and the subsidence of the western Salt Range region
relative to the eastern Salt Range during the Mesozoic attest to pre-
orogenic movements. Magnetic-polarity stratigraphy indicates that
since the onset of the Himalayan orogeny, the Potwar Plateau was the
locus of molasse sedimentation; it apparently subsided rapidly during
early Siwalik deposition (approx. 15-5 mya). Since then, the regions
to both the west (Trans Indus Salt Range) and east (Jhelum Reentrant)
became the locus of sedimentation in the foredeep (fig. 2; Khan, 1984).
Today, the Jhelum Reentrant remains a structural trough subsiding more
rapidly than adjacent parts of the molasse basin (ie. Potwar Plateau;
Johnson et al., 1979).

Methods

The two main tools used in the structural analysis of thrust belts
are balanced cross-sections (Dahlstrom, 1969) and the interconnections
or relationships among faults (Boyer and Elliot, 1982). The intercon-
nections of faults are studied qualitatively using observations of
branch and tip lines, for example, to construct branch line maps, while
balanced sections require a more analytical approach. For this reason,
only the techniques involved in section balancing, including restoring
sections, are explained here. Section balancing, including the deter-
mination of shortening, is subdivided into the following categories:
(1) conversion of seismic time sections to depth sections; (2) reinter-
pretation of the depth sections to balanced sections, incorporating
surface and drilling data; (3) restoration of the balanced sections;
and (4) shortening calculations.
Depth Sections: Surface constraints on the interpretation of the seis-
mic reflection transects (fig. 3) were the geologic maps of the Salt
Range (1:50,000 scale) by Gee (1980), and maps by Cotter (1933) and
Gill (1952) for the Potwar Plateau. Faults and fold axes identified on
Landsat imagery (approx. 1:250,000 scale) were used to supplement geo-
logic maps in the northern Potwar Plateau. Strikes and dips, corrected
10

to apparent dips, were projected along strike into the line of section
from a couple of kilometers away in the frontal sections to up to five
kilometers away in the Potwar Plateau. Well information (fig. 3; see
Leathers [1987] for additional well data), particularily formation
thicknesses, were also projected onto sections paying particular atten-
tion to the depositional strikes of units. An additional subsurface
constraint was the structure of the basement surface, as determined via
gravity modeling by Duroy (1986; Duroy et al., in press).
Velocities used for depth conversions (e.g. fig. 4) were obtained
by incorporating: (1) the range of possible velocities for a given
lithology (Telford, 1976); (2) stacking velocities from seismic lines;
and (3) the velocity structure calculated by Leathers (1987) from sonic
logs. The velocity structure was fine tuned on the frontal sections by
modeling velocity effects using the AIMS (trademark Geoquest, Int.)
program, including automatically migrating sections during depth con-
version (fig. 4). The main depth section was extended to the north
from the AIMS sections by hand calculations to depth using the velocity
constraints listed above.
Balanced Sections: Frontal section lines were chosen along seismic
lines (fig. 4), which are slightly oblique to the transport direction,
while the main balanced section was constructed parallel to the trans-
port direction of the Salt Range thrust, taken as S 15° E (Yeats et
al., 1984). Vertical, rather than downplunge sections were construct-
ed, because major structures along the transects have negligible
plunge. Data were assembled by drawing topographic profiles and then
plotting the outcrop constraints. Next, the geometry of the sections
were modified by extrapolating and interpolating via the Busk and kink
methods. The Busk method assumes concentric folding whereby bed con-
tacts are drawn between inflection points around their centers of
curvature. The kink method assumes straight limbs and angular hinges,
but can be used to approximate all curved shapes. Axial surfaces
bisect the angles between fold limbs and in this way the kink method
has been used to assure a retrodeformable section (Suppe, 1985).
The sinuous bed method, which requires that beds have a constant
length throughout the stratigraphic section so that beds extend into
their original stratigraphic wedge when retrodeformed (Dahlstrom,
11

Fig. 4: Example of velocities used for seismic time to depth conver-


sion (i.e. for section C-C1). (A) Digitized time section showing
velocities in m/sec. (B) Cross-section after depth conversion,
but prior to balancing. During cross-section balancing, velocity
effects are smoothed out and details added from surface maps and
drilling information. Formation abbreviations: T-QM=Tertiary to
Quaternary molasse; MS=Middle Siwaliks; DP=Dhok Pathan; Na=Nagri;
LS= Lower Siwaliks; Ch=Chinji; RM=Rawalpindi Group (molasse);
e-E=Cambrian to Eocene (i.e. platform strata); SRF=Salt Range
Formation; pe=Precambrian (crystaline) basement.
10km 20km 30km
i r 1 0
2400
TOPOGRAPHIC SURFACE
,..
2900
310 0
a
3100
z
2o
3500 3 600 U
360 0 W
4 500
co
4500 3

6000 6000
A
.
1
4

0 20km 30km
1 1
TOPOGRAPHIC SURFACE
SEA SEA
LEVEL MS (DF:No) LEVEL
u)
CC
LLI

2 2 1w
2
0
4 4 -.1
;2

Figure 4
13

1969), was used to check the viablity of all balanced sections (see
Elliot, 1983). Sections were constructed at a scale of 1:78,740.
Because a single bed length measurement does not constitute section
balancing, eroded beds were restored so that consistency of bed length
could be used to check the validity of interpreted structural styles.
Measured bed length variations were held to under 1%, with the excep-
tion of the NPDZ portion of the main balanced section which had a 1.3%
error.
Due to ductile flow within the Salt Range Formation, only the
overlying horizons could be checked by measuring bed lengths. Follow-
ing the assumption of conservation of area in cross-sectional view
(Dahlstrom, 1969), the Salt Range Formation was area balanced (e.g.
Gwinn, 1964), to correspond to a restored model, on the main balanced
section using a planimeter. Keuffel and Esser Co.'s compensating polar
planimeter (model # 62 0000) could be used to measure areas to within
0.01 sq. in., which is the equivalent of 0.04 sq. km. on the cross-sec-
tion. Area balancing of the Salt Range Formation assumed that equal
amounts of salt entered and left the section during deformation. Salt

areas originally north and south of the basement offset were balanced
separately, thereby supplying a constraint on the original thickness of
the Salt Range Formation at the basement offset (see restoration
methods and appendix B).
A pair of reference lines, called pinlines, marking the ends of
balanced sections, were chosen in areas of little or no interstratal
slip. The undeformed foreland was chosen as the frontal pinline on all
balanced sections. The rear pinlines were placed along the axis of the
Baun syncline on the frontal sections (see sections A to E, fig. 3) and
near the rear of a thrust slice in the NPDZ for the main balanced
section (F-F', fig. 5). In addition, local pinlines were used within
the main section, at the fold axes of the Baun and Soan synclines and
just south of the Salt Range frontal thrust, to further constrain the
shortening across individual structures (e.g. Woodward and Boyer,
1985).
The most important criterion in constructing a retrodeformable
section across a faulted terrane is matching hanging-wall and footwall
cutoffs. In this way footwalls exposed at the surface reflect the
14

Fig. 5: Structure map of the SR/PP including locations of paleomag-


netic sample sites (with paleomagnetic pole directions where
available) and other figures illustrated in text. Solid structure
symbols are more pronounced features, mapped from LANDSAT photos.
Dashed symbols denote less prominent structures whose locations
were derived from seismic, surface maps, and well control. Dotted
pattern represents buried structures. Structure abbreviations:
Ba=Balkassar Anticline; BF=Basement Fault; BS=Baun Syncline;
Bu=Buttar Anticline; C=Chambal Ridge; CF=Chhumbi Fault; D=Dhulian
Anticline; DR=Domeli Ridge; HR=Hill Ranges; J=Jabbar Anticline;
JM=Joya Mair Anticline; JT=Jogi Tilla Anticline; K=Khaur Anti-
cline; KDF=Karangal/Diljabba Fault; KF=Kalabagh Fault; KK=Kotal
Kund Syncline; KKF=Kallar Kahar Fault; L=Lilla Anticline; Le=Lehri
Anticline; M=Mahesian Anticline; MA=Meyal Anticline; MBT=Main
Boundary Thrust; MF=Murree Thrust; 0=Oblique Ramp; P=Pabbi Hills
Anticline; R=Rhotas Anticline; RF=Riwat Fault. Paleomagnetic
sample locations: (1) Andar Kas; (2) Baun; (3) Bhangala; (4) Dala
Nala; (5) Dina; (6) Ganda Paik; (7) Hasal Kas; (8) Jalapur;
(9) Kamliel Kas; (10) Kas Dovac; (11) Khaur Composite; (12) Kotal
Kund; (13) Nagri; (14) Pabbi Hills; (15) Rada Kas; (16) Rhotas;
(17) Soan; (18) Tatrot. Paleomagnetic sites adapted from Keller
and others (1977), Barndt and others (1978), Opdyke and others
(1979), Tauxe and Opdyke (1982), Johnson and others (1982), Opdyke
and others (1982), Raynolds and Johnson (1985), and Burbank and
others (in press).
15

730
72 0

33

330

SALT RANGE FRONTALTH


L

730
0 50 k m
..),....y/ THRUST FAULT; TEETH ANT !CLINE PALEOMAGNETIC
ON UPTHROW BLOCK ti) SAMPLE SITE
,...Y NORMAL FAULT; TICK ......15/ SYNCLINE r........, DIRECTION OF PALEO-
MARKS ON DOWNTHROWN BLOCKS MAGNETIC POLE

Figure 5
16

shape of eroded hanging-walls, and hanging-walls seen at the surface


represent the shape of burried footwalls. However, a special problem
exists when attempting to balance a thrust sheet across a ramp. Slip

may not be conserved on cutoffs along the thrust fault. This variation

in slip is currently a point of controversy (see Woodward and Boyer,


1985). Slip is thought to increase upwards on a thrust fault (Elliot,
1976), or thrusts lose displacement upwards (Fain, 1967; Dahlstrom,
1969; Thompson, 1981; Suppe, 1983) as a thrust sheet overrides a ramp.
The selection of pinline positions north of the ramp, within the Baun
syncline, followed Suppe's (1983) assumption that there is no inter-
stratal slip at the trailing edge of a fault-bend fold. Therefore,

shear due to ramp climb was transferred to the leading limb. As a

result, shear steepened the hanging-wall ramp so that hanging and foot-
wall ramp shapes differ slightly. If, alternatively, the footwall and
hanging-wall ramps were made to match, shear would impart interstratal
slip on the back limb (i.e. the remainder of the thrust sheet extending
to the northern Potwar Plateau; Woodward and Boyer, 1985).

Restorations: In order to restore a section to its undeformed state,


the original state of the strata must be known or assumed. In thrust

faulted terranes, the basement or strata beneath the decollement are


held fixed, and fault displacements are reversed by moving the hang-
ing-wall rocks along the thrust fault surface. However, in the present

study area there are two major problems. First of all, the salt hori-

zon has flowed, and therefore its original depositional geometry is


unknown. Secondly, the lower plate has been flexed downward to the
north by an unknown amount.
Retrodeformation was accomplished by restoring bed thicknesses,
starting at the base of the sedimentary section, relative to the
sinuous bed length distance from the frontal pin line. In areas of

nonparallel folding near the Salt Range front, undulations were


smoothed and the average bed thickness across the entire fold was used.
The earliest restoration was performed first. For the later restora-

tion, the basement and overlying sediments were flexed by the desired
amount and additional sediments stacked above.
Because molasse deposition was contemporary with thrusting,
restoration required knowing the precise ages of sediments and timing
17

of deformation events. Individual mapped formations in the Pakistan


foreland are time transgressive, therefore formations cannot simply be
stripped away. For these reasons, convenient times of restoration were
chosen such that intervals of strata, representing distinct assemblages
of lithologies whose ages are constrained, could be sequentially
removed. These restoration times correspond to post-Eocene/pre-Murree
deposition (approx. 35 mya), and post-Middle Siwalik/pre-Upper Siwalik
deposition (approx. 5 mya; see F-F' fig. 5). In consideration of the
previously mentioned complexities and the times of restoration, the
following assumptions were made: (1) salt originally north and south of
the basement offset could be (area) balanced independently; (2) for
the 35 mya restoration, the top of Eocene north of the basement offset
was horizontal; (3) south of the basement offset, post depositional
erosion of Eocene and older strata may have occurred; and (4) a topo-
graphic (i.e. depositional) slope existed towards the axis of the
foreland basin during molasse deposition. Assumptions (2) and (4) aid
in estimating basement dips at the respective times of restoration. In

this way, the restoration process itself geometrically constrains the


amount of flexure. Appendix B discusses the results and implications
of these assumptions and the retrodeformation.
Finally, in a deviation from standard restoring techniques, Indian
crust along with its sedimentary cover, was pulled back through time
according to its present flexed shape (e.g. Duroy, 1986; Duroy et al.,
in press). Because the shape of flexural bulges can vary with time, a
generalized flexure model was assumed. From south to north, this model
consists of an area of flat basement corresponding to the crest of the
bulge (currently the Sargodha High), steep basement dip on the flank of
the bulge, and then a zone of shallow to moderate basement dip (cur-
rently the Jhelum River plain and Salt Range) giving way to steep
basement dip in the area of maximum flexure (currently the Potwar
Plateau and Hill Ranges).
Calculation of shortening: Shortening was calculated by subtracting the
present horizontal distance from the deformed bed lengths (Dahlstrom,
1969). These values differ slightly from those obtained by subtracting
the straight-line distance between pinline cutoffs (Suppe, 1985), or
the basement length, because beds were tilted prior to thrusting.
18

Subtracting the horizontal distance was preferred because reversal of


interpreted extension during the basement faulting event is not taken
into consideration. The results are also directly comparable to plate

motions. True shortening in the horizontal plane should also use the
horizontal projection of the original bed length. However, errors
introduced in this study by pre-deformational tilted beds are at most
0.1 km, or about 0.5%. Shortening values were corrected for obliquity
using a cosine correction, which is appropriate for the low angles
involved here (Cooper, 1983).
Experimental error was minimized by using the same instruments and
technique to determine both the deformed and original bed lengths.
Uncertainties such as those derived from imprecise measurements, bed
tapers, and most other factors (see appendix C) are covered by the
arbitrarily assigned 5% error. An exception is the balanced section
across the NPDZ (see F-F', fig. 5), where only a moderate degree of
confidence is assigned to its structural interpretation; estimates of
shortening there, based on deformation style and not precise inter-
pretations of individual structures, have been arbitrarily assigned a
33% error based on these uncertainties.
19

STRATIGRAPHY AND SEDIMENTATION

Stratigraphy

The basement beneath the SR/PP is probably a northeastern exten-


sion of the Indian craton. Precambrian metasediments and plutonic
rocks constitute the most northernly exposure of the Indian Shield at
the Kirana Hills (Sargodha High; fig. 2). Resting directly on these
crystalline rocks north of the ridge, the Salt Range Formation or
"saline series", consisting of mainly gypsum, salt marls and rock salt,
has been correlated with the Hormuz Formation of the Persion Gulf
(Zarkov, 1981). Both Precambrian and Cambrian ages have been assigned
to the Salt Range Formation by various workers; therefore, its age is

referred to as Eocambrian. Existence of evaporites in the Hazara


Slate Series suggests that the Salt Range Formation may extend north-
ward from its outcrops at the Salt Range at least as far north as
Hazara (Latif, 1973). This is a significant hypothesis for the sub-
Himalaya of Pakistan, because the evaporites form the zone of decolle-
ment. Nonmarine to shallow marine sandstones and shales of the Cam-
brian Jhelum group overlie the Salt Range Formation, except in the
western Salt Range area. Similarity of early Paleozoic stratigraphy
throughout southwestern Asia implies the developement of a widespread
Gondwanan passive continental margin following late Precambrian orogen-
esis (StOcklin, 1974; Bender, 1975; Cherven, 1986).
Above the first of three major unconformities, lower Permian
strata overlap the Cambrian westward. The Gondwanan Tobra Formation,
at the base of the Permian succession, contains distinctive Precambrian
granitic clasts from the Indian shield. Marine influence increases as
clastics are replaced by calcareous rocks in the upper Permian. A
Mesozoic sequence of predominantly marine clastics overlies the Permian
in the western half of the Salt Range. The strike of Mesozoic units
prior to deposition of overlying rocks in the Potwar Plateau suggests a
west-northwest facing, Mesozoic passive margin (Yeats and Hussain, in
press). The next major unconformity is at the base of the Tertiary,
where marine strata of Paleocene and Eocene age overlap Mesozoic and
Permian rocks eastward. Basin analysis indicates that Paleogene strata
20

were deposited in a basin, centered to the north near the Kala Chitta
Range (Wells, 1984).
Above the last of the major unconformities are fine-grained
clastics of the Murree Formation, derived from the Himalaya. The

Murree Formation, along with the lower Siwalik Kamlial Formation make
up the Rawalpindi Group. Succeeding the Murrees across the Potwar
Plateau are massive sandstones and clays of the upper Miocene to
Pliocene Siwalik molasse. Confusion abounds concerning classification
and subdivision of the Siwaliks. Early workers attempted to subdivide
the Siwaliks using fossil assemblages; however, due to the scarcity of
fossils it became more useful to subdivide the Siwaliks into litho-
stratigraphic units. This is the present system of classification.
Pilgrim (1910) subdivided the Siwalik Group into the Upper, Middle and
Lower Siwalik subgroups. In 1913, Pilgrim subsequently subdivided
these subgroups into formations; this classification remains today,
with only minor changes. The Lower Siwaliks correspond to the Kamlial
and Chinji Formations, while the Middle Siwaliks are divided into the
the Nagri and Dhok Pathan Formations. Facies changes, and therefore
formation contacts, are time transgressive. Sandstones within the Soan
Formation (Upper Siwalik) grade laterally into boulder conglomerates,
indicating the encroachment of Himalayan thrust sheets. Reflection
data show that the undeformed molasse section is as thick as 5500 m in
the Soan syncline.
Post Siwalik deposits are critical for dating deformation; unfor-
tunately much of the younger strata have been removed by uplift and
erosion. An exception is the Lei conglomerate preserved in the Soan
syncline south of Rawalpindi, which has a basal age of about 1.9 m.y.

(Raynolds, 1980). The Lei conglomerate, like the Kalabagh conglomerate


of the southwestern Salt Range, is a valley fill conglomerate with
Eocene clasts. Post Siwalik conglomerates are related by similarity of
source rocks and mode of formation, while their ages are related to
local events which controlled their deposition. At the top of the
section in portions of the Potwar Plateau is the Potwar Silt. The age

of this formation (see appendix A [Baker et al., in review]) is


significant because it is believed to have been ponded behind the
rising Salt Range (Yeats et al., 1984). Recent work has shown that the
21

Potwar Silt is a loess derived from an easterly source (Shroder, in


prep).

Sedimentation and Depositional Models

As a result of continent-continent collision, a structural trough,


called the Indo-Gangetic foredeep was formed on the Indian plate.
Coarse-grained alluvial-fan sediments were deposited along the basin
margin (proximal), while finer grained fluvial sediments were deposited
downstream (medial and distal). In response to continued northward
movement of the Indian crust, the foredeep migrated relatively south-
ward across the foreland.
In the Indian foreland, Acharyya and Ray (1982) documented that
the southward progression of the thrust front is accompanied by
migration of the Siwalik depositional axis. The alluvial plain there
can be subdivided into a northern domain where rivers flow south from
the mountains across broad distributary fans, a central area where
south-flowing rivers merge into the Ganges and flow parallel to the
Himalayan front, and a southern domain where smaller rivers flow
northward from the Indian shield. Ideally, sediments deposited in
these three domains are juxtaposed by the southward migrating molasse
basin, such that the molasse sequence is underlain by shield derived
alluvial sands and topped by Himalayan fan gravels.
In Pakistan, where the Siwalik molasse outcrop is broader and has
been age-calibrated more extensively by magnetostratigraphy and tephr-
achronology, a more detailed tracking of facies migration has been
documented. Distal, white sand sandstones deposited by a major trunk
stream with east-northeasterly paleocurrents (Raynolds, 1981) are
replaced upsection by brown sandstones with a nearby Himalayan prove-
nance (medial) and south flowing paleocurrents. The lithologies of the
white and brown sandstones resemble those which are carried by the
modern Indus and Jhelum rivers respectively (Burbank and Raynolds, in
press). In the Jhelum reentrant, the observed coarsening-upwards of
the Himalayan molasse, due to outward displacement of the orogenic
front, is capped by the Soan boulder conglomerate. By measuring the
southward younging of the first appearance of this 3 to 1.5 m.y.
22

boulder bed, Raynolds and Johnson (1985) estimate that the rate of
southward progradation of lateral facies has been up to 30 mm/yr. An
exception to the ideal foredeep migration model, in which facies
contacts would be time-transgressive to the south, is the Nagri-Dhok
Pathan contact (approx. 7.9 to 8.6 m.y.) which is northward time-trans-
gressive in the central Potwar Plateau (Behrensmeyer and Tauxe, 1982;
see Leathers, 1987).
Numerous magnetostratigraphic sections have been measured across
the SR/PP and within the Jhelum reentrant (located on fig. 5). By

correlating these sections to each other and to a magnetic reversal


time scale (e.g. Cox, 1969), sedimentation rates for different time
intervals can be derived. At most locations plots of accumulated
sediment thickness versus time show low initial sedimentation rates in
a distal foredeep, followed by maximum sediment accumulation rates in
the foredeep depocenter, and finally an attenuation of sedimentation
rates due to the encroachment of deformation. Sedimentation ceases as
the area is deformed and positive relief attained. The resulting
sigmoidal plots suggest that the depocenter has migrated southward at
over 20 mm/yr (Raynolds and Johnson, 1985).
When a thrust of regional extent ramps toward the foredeep basin,
a new intermontane basin may form on the back of the thrust sheet.
Behind the thrust, pre-existing fluvial systems may be ponded. Burbank
(1982) suggests that the Peshawar, Campbellpore and Kashmir basins were
formed in this way (fig. 2). These intermontane basins are probably
ephemeral features whose sediments will most likely be removed by later
stages of deformation (Burbank, 1983). In the Potwar Plateau, the
Potwar silt may represent an intermontane basin deposit ponded by the
rising Salt Range. The Potwar silt is currently being dissected and
within a few tens of thousands of years will probably be completely
eroded (R. Yeats and R. Lillie, writ. comm., 1984).

Depotectonics

Depocenter Migration across the Central SR/PP: Depositional models


based on Upper Siwalik sediments in the Jhelum reentrant (e.g. Raynolds
and Johnson, 1985) may not apply to older molasse deposition across the
23

central SR/PP. For example, the southward shift of isochron thickness


maxima observed in the Jhelum Reentrant is not found across the Potwar
Plateau (see Leathers, 1987); in fact, local northward progradation has
been documented (Behrensmeyer and Tauxe, 1982). In addition, the Lower
and Middle Siwaliks (Chinji through Dhok Pathan), thin gradually across
the Potwar Plateau, and then appear to thin abruptly farther towards
the south across the Salt Range. The Toot #2 well in the western
Potwar Plateau penetrated 2952 m, while 2510 m of Lower and Middle
Siwaliks were drilled by the Kot Sarang and Karsal wells in the central
Potwar Plateau. Only 1425 m of molasse was penetrated by the Lilla
well (see fig. 3 for well locations). The total thickness of the
Chinji through Dhok Pathan molasse is estimated to be 3100 m in the
Soan syncline and 2700 m in the Baun syncline (fig. 6). South of the
central Salt Range the entire molasse section is less than 2 km;
however, most and possibly all of this molasse is interpreted to be
young (i.e. Upper and Post Siwaliks). It is proposed that depocenter
position was relatively stationary during the deposition of Lower and
Middle Siwaliks from about 15-5 mya (Johnson et al., 1985; Johnson et
al., 1982). Basin subsidence and therefore the locus of sedimentation
may have been fixed by downdropping along the basement normal fault
(fig. 6). Since this time, the depocenter jumped south of the Salt
Range. The rise of the Salt Range may have caused a reorganization of
drainage patterns (e.g. Burbank and Raynolds, in press), resulting in
the shift in depocenter positions.
ForedeeR Migration Rates: Foredeep position is controlled by the
location of the lithospheric load; similarly, foredeep migration is
influenced by movement of the load. Because the Himalaya functions as
the main lithospheric load, Molnar and Deng's (1984) slip rate of 18
mm/yr for underthrusting the Himalaya in India is an important indica-
tor of foredeep migration; it approximates the northward movement of
Indian crust relative to the Himalaya. Direct evidence of the foredeep
migration rate was obtained by Lyon-Caen and Molnar (1985), who cali-
brated the southward younging of basal molasse sediments overlying the
Indian shield at 10-15 mm/yr in India.
The above rates are, however, significantly less than the facies
and depocenter migration rates determined by Raynolds and Johnson
24

Fig. 6: Schematic diagrams illustrating the general southward shift of


molasse deposition. (A) Near the end of distal Rawalpindi molasse
deposition across the Potwar Plateau (approx. 15 mya). Assuming
steady state basin shape, there is a problem with restoring old
molasse in the foredeep using derived migration rates (discussed
in text). Molasse above jagged line eroded as area passed across
peripheral bulge (implying sediment bypass as bulge emerged).
Alternatively, the peripheral bulge did not exist or the foredeep
basin was much wider. If so, the basement would have continued
its north dip and Rawalpindi molasse lapped against the edge of
the basin. (B) At the end of Middle Siwalik molasse (Dhok Pathan)
deposition (approx. 5 mya). Lower and Middle Siwalik molasse
deposition was concentrated in the Potwar Plateau attaining its
maximum preserved thickness in the Soan syncline. The depocenter
may have been confined in this region by basin downdropping along
the basement normal fault. Lower and Middle Siwaliks thin abrupt-
ly south of the Baun syncline across the basement offset. The
bulk of the sediments was derived from the deformation front
located somewhere to the north; however, evidence suggests
sporadic erosion and transport from the crest of the peripheral
bulge. (C) Present situation. Upper and post Siwalik molasse
deposition was concentrated south of the Salt Range and to the
east in the Jhelum Reentrant. The bulk of the sediments south of
the Salt Range are young, because this region has only recently
subsided into the foredeep. The present depocenter for sediments
derived from the Salt Range and elsewhere is located in the Jhelum
River Plain.
HINTERLAND
SOURCE UNPRESERVED
PROXIMAL POT WAR PLATEAU
MOLASSE PRESENT TOP OF
DISTAL MOLASSE DEPOS ITION RAWALPINDI MOLASSE

EMERGI G PERIPHERAL: BULGE .

BAUN SYN.

.A UPPER-POST SIWALIKS CAMBRIAN TO EOCENE


° o : AND QAL
LOWER- MIDDLE TERTIARY EOCAMBRIAN SALT RANGE
SIWALIKS MOLLASSE FORMATION

RAWALPINDI GROUP PRECAMBRIAN BASEMENT


Figure 6
26

Fig. 7: Restored positions of the Sargodha High, Salt Range Frontal


Thrust, basement fault, and Baun locality at 4.5 mya, holding
Indian crust fixed. The Salt Range thrust sheet has been pushed
back 25 km, N15°W. The Sargodha High is assumed to be a peri-
pheral bulge through which Indian crust has migrated northward,
controlled by plate convergence directions (see Minster and Jordan
[1978] for pole of rotation). The restoration direction of the
bulge is not controlled by plate convergence directions, but
rather by the shift of the lithospheric load (Himalaya) versus the
Indian crust. In other words, the relative slip direction between
the Indian crust and the western and central Himalaya in northern
Pakistan, India, Nepal and Tibet controls the direction of bulge
migration. The slip direction and therefore bulge migration is
assumed to have been perpendicular to the northwest-southeast
trend of the Himalaya and Sargodha High. The minimum rate of
movement between the bulge and Indian crust is constrained by the
derived frontal shortening rate, while the maximum rate was taken
as Molnar and Deng's (1984) estimated slip rate beneath the Hima-
laya (18 mm/yr). The illustration depicts collision of the Sar-
godha High and the western Salt Range, resulting in thrust ramping
(Leathers, 1987). -Note that the basement fault restores to the
northern flank of the peripheral bulge. Dark areas along the
Sargodha High represent exposures of Precambrian rocks forming
the Kirana Hills.
RESTORED POSITION
OF SALT RANGE
FRONTAL THRUST
RESTORED POSITION
OF BAUN SITE
25 SALT RANGE
BASEMENT BAUN SITE
k FRONTAL THRUST
OFFSET
POSSIBLE RESTORED
POSITIONS OF PERIPHERAL
BULGE AXIS

18mm/yr
28

(1985), who suggest that there is a rough equilibrium between plate


convergence rates and facies/depocenter migration. Instead, the high
progradation rates were influenced by increased source area uplift
rates and thrust sheet advance (Yeats, in prep). The encountering of a
salt-lubricated decollement causing a rapid migration of the deforma-
tion front may be responsible for the higher rates obtained by Raynolds
and Johnson (1985).
There is also a problem with restoring Lower Siwalik and Murree
molasse in the narrow foredeep using derived rates of foredeep and
peripheral bulge migration (e.g. fig. 7). Even at the lowest rate of
10 mm/yr, assuming a steady state foredeep basin shape, these sediments
restore south of the peripheral bulge (Sargodha high; fig. 2). Obvi-
ously there is a problem with the age of Himalayan molasse in the
present, narrow Pakistani foredeep. There are several explanations
for this paradox: (1) the foredeep basin was broader in the past
because either the Sargodha high is a remnant structure and moves with
the Indian plate, or the foredeep is narrowing due to bulge migration
towards the thrust front; (2) foreland basin deposition was not re-
stricted by a peripheral bulge because sediments bypassed the peripher-
al bulge or the peripheral bulge did not exist (fig. 6); and (3) the
relative rate of movement between Indian crust and the peripheral bulge
was slower than presently indicated, as would have occurred if a larger
amount of convergence between the Indian plate and Eurasia were absorb-
ed north of the collision zone (see Molnar et al., 1987) in the past.
29

STRUCTURE OF THE CENTRAL SALT RANGE AND POTWAR PLATEAU

General Description

The Salt Range and Potwar Plateau are part of a 200 km wide
salient protruding southward between the Mianwali and Jhelum Reentrants
(fig. 2). Approaching the area from the south, the Salt Range escarp-
ment appears to rise abruptly from the Punjab plains. To the north,

the Potwar Plateau is bordered by the Kala Chitta and Margala Ranges,
which are collectively referred to as the Hill Ranges. In the Hill
Ranges, intensely deformed, Precambrian slates plus Mesozoic and Tert-
iary strata are exposed at the surface along a series of south directed
thrusts. This fault zone is believed to be a westward extension of the
MBT of India (figs. 1 and 2; Yeats and Lawrence, 1984).
The Salt Range is an east-northeast trending highland, bent north-
ward at both ends. The thrust front at the Salt Range is strongly
emergent (terminology of Morley, 1986). To the west, it is offset from

the weakly emergent thrust front at the Surghar Range by the right-lat-
eral Kalabagh tear fault (fig. 2; McDougall, 1985). To the east the

frontal thrust becomes blind and the Salt Range dies out into the
Chambal Ridge and Rohtas anticline (Yeats et al., 1984). The Pabbi
Hills anticline (fig. 5) appears to be the eastward extension of the
deformation front across the Jhelum River.
The Potwar Plateau is an elevated area of little relief, sometimes
referred to as the Potwar trough (e.g. Voskresenskiy, 1979) or Potwar
depression (e.g. Khan et al., 1986), due to its thick sedimentary fill.
The Potwar Plateau can be divided into the NPDZ, which merges along
strike with the Kohat Plateau to the west, and a southern area of
little deformation, called the Soan trough because of its alignment
with the Soan River.
The SR/PP is dominated by the Salt Range thrust sheet. It extends

from a branch line beneath the northern limb of the Soan syncline
(fig. 8) to the Salt Range front, terminated to the west by the Kala-
bagh fault and to the east by the Karangal/Diljabba fault (KDF) and
a diffuse zone of deformation in the eastern Potwar Plateau. Overall,

the shape of the Salt Range thrust sheet, presence of oblique folds
30

Fig. 8: Block diagram illustrating change in deformation style across


oblique ramp. (A) Duplexing east of the oblique ramp and imbri-
cation west of it are consistent with surface geology. Surface
geology adapted from Wadia (1928) and Cotter (1933). Central
portion of map interpreted because geologic contacts and faults,
on bordering published maps, do not match. (B) Cross-section
(G-G' in A) of oblique ramp marking the westward termination of
NPDZ frontal duplexing. Note that the roof thrust dips east-
wards. A transect from the syncline (see Pariwali well in [A]) to
G' would show more typical oblique ramping or a culmination wall
(e.g. Butler, 1982). Portion of a branch line map (e.g. Hossack,
1983) shown at surface. Branch lines denoted by broken lines.
Right side (with semicircles) = trailing branch line of the Salt
Range thrust sheet, left side (with semicircles) = lateral/trail-
ing branch line of NPDZ frontal horse, and left side (with teeth)
= leading branch line of NPDZ frontal horse. Branch lines picked
at base of platform sequence. Note trailing edge of Salt Range
thrust sheet in (A) and (B). Rawalpindi Group is divided into the
Kamlial (K1) and Murree (Mu) Formations. Other abbreviations are
the same as in figure 4.
31

WSW ENE

:7;:Vf,(:ViNs,=12=^0xFilAt;"4st,.4;11:11:19
44Ass4^:tr
,.. .01 lb l"!

ASMANNNO4r

V.E..1:1
10 km
Figure 8 (cont.)
32

near its lateral terminations and concentration of layer parallel


shortening (via salt growth) near the range front resemble Fischer and
Coward's (1983) model of a surge zone with compressional flow.

Northern Potwar Deformed Zone

Description: South of the Hill Ranges, folds and thrusts expose Eocene
and younger strata in the NPDZ. Surface structure consists of a folded
zone with anticlines, isoclinally overturned to the south, giving way
to thrusting farther south. In the southwest, map patterns of Murree
and Siwalik strata depict large anticlines and synclines cut by thrust
faults (see Cotter, 1933). To the north and east, only Murrees are
exposed, making folds and thrusts difficult to map; however, some large
folds are visible on Landsat photos. Along the southern margin of the
NPDZ (i.e. steeply upturned north limb of the Soan syncline), published
maps show four major, south-verging thrust faults that merge eastward
into one (Gill, 1952; Voskresenskiy, 1978). The existence of other
faults is suggested in the detailed mapping of the Dhulian, Khaur, and
Meyal-Kharpa anticlines (fig. 5) by Martin (1962).
The complex seismic expression of overthrust strata in the NPDZ is
indicative of the intense folding and thrusting (figs. 9 and 10).
Detailed interpretations are not possible; instead, observations are
made from which structural style can be inferred. Interpretations
of lateral changes in structural style are hindered by the lack of
strike lines and the wide spacing (5-10 km) of available seismic tran-
sects (fig. 3).
On most lines the easily discernible basement reflections of the
Soan syncline can be traced northward beneath the NPDZ (fig. 11).
Farther north, basement reflections become more difficult to identify
(fig. 10). The discontinuous, subhorizontal reflections of the base-
ment merge with platform sequence reflections, as the transparent zone
corresponding to Eocambrian strata thins (see Lillie et al., 1987;
their fig. 4). The platform sequence is deformed into concave downward
arcs and "s" shaped features, interpreted to represent thrusted anti-
clines and horses (packages of rock bounded by thrust faults; figs. 9
and 10; Boyer and Elliot, 1983). The disharmonic nature of these
33

Fig. 9: Uninterpreted (A) and interpreted (B), unmigrated seismic line.


Note disharmonic folding where syncline (beneath Mianwali well)
overlies anticline. The lower (Rawalpindi) molasse appears
to be tectonically thickened beneath the syncline (shown schema-
tically by arrows beneath syncline). No attempt was made to
interpret the connections between mapped surface thrusts and
subsurface thrusting. Mianwali well penetrated 2116 m of Siwal-
iks, 2733 m of Rawalpindi and 166 m of platform strata. Arrows at
the top of reflection profiles mark the positions of mapped sur-
face thrusts. See figure 3 for location. TM=Tertiary molasse;
SM=Siwalik molasse, other abbreviations are the same as in figure
4.
rn
-
35

Fig. 10: Uninterpreted (A) and interpreted (B), unmigrated seismic


transect showing reflections within the "triangle zone", near the
southern terminus of the NPDZ. Reflections may represent stacked
horses of platform strata although antiforms are tighter and
smaller than they appear on this unmigrated section. The decolle-
ment is probably just above the 4 second line. Dhurnal wells
penetrated an average of 2170 m of Siwaliks, 1560 m of Murrees,
and 1120 m of platform strata before penetrating the Salt Range
Formation. R=roof thrust (upper-level detachment) other
abbreviations and symbols are the same as in figure 9. See
figure 3 for location.
ch SECONDS
:0+5+-I
mm (A n.) 0
- . - -
vIz:t,N .,,J z
-I'tic IfitlAlike...-' 4-5,v---,":3%.::44q-...:";'':

r is?

CD

m
ro c
V!
3

.s 11j tn

m Im
37

Fig. 11: Uninterpreted (A) and interpreted (B), unmigrated seismic


line across the northern limb of the Soan syncline. Note
thickened lower molasse section interpreted to be the result of
back thrusting along an upperlevel bedding plane detachment or
roof thrust ("R"). Reflections become chaotic on the left (north)
end of the section, towards the center of the "triangle zone".
For this reason, the northward extent (i.e. footwall cutoff) of
platform strata is unknown. Platform strata may also be present
between the upper two thrusts. Abbreviations are the same as in
figure 9. See figure 3 for location.
38

0 5km

:
4A.N.

Figure 11
39

reflections, compared to the subhorizontal ones below, support the


interpretation of an intervening decollement. Overall, the seismic
signature becomes less discernable to the north, in accordance with the
increased intensity of deformation in that direction (left side fig.

10).
An oblique ramp marks a lateral change in deformation style
(fig. 12). The change in structural style is shown at the surface by
coalescing thrust faults and in the subsurface by different seismic
signatures (compare figs. 9 and 10). Farther south, the change in
deformation style is marked at the surface by the curved axis of the
Khaur anticline (fig. 5), where the oblique ramp represents the western
edge of horses involved in duplexing beneath the northern limb of the
Soan syncline (fig. 8).
West of the oblique ramp, several lines show the stacking of
reflectors into an apparent imbricate fan. Surface maps indicate that
thrusts cut progressively younger rocks to the south (see Cotter,
1933), suggesting an imbricate fan involving the entire allochthonous
section. On one of the few migrated lines, a series of five to seven
"s" shaped horses form a pattern resembling a hinterland-dipping duplex
(see Boyer and Elliot, 1983). In an attempt to interpret the inter-
connections of faults, two observations are made. First, faults
interpreted at depth do not always appear to connect with thrusts
mapped at the surface. Second, there are more thrusts at depth than
mapped at the surface (figs. 9 and 10). The Mianwali well (figs. 3 and
9) penetrated almost twice the normal thickness of the Rawalpindi
Group, suggesting that a thrust cuts at a low angle through this
formation. A bedding-plane detachment may therefore be present
within the Rawalpindi Group, possibly near its base. Furthermore,
evidence of disharmonic folding between the reflective platform
sequence and near-surface reflections is seen on several seismic lines
(fig. 9).
East of the oblique ramp, seismic observations are limited, due to
a general decrease in data quality. Two important patterns of reflec-
tions are noted near the southern margin of the deformed zone. One is

the triangular pattern formed by the south-dipping, north limb of the


Soan syncline (fig. 11) in combination with the north dipping reflec-
7 2° 30'
HILL RANGES
44/RREE FAULT (A4
10km
33°30' HINTERLAND

/ DIPPING
N
33°30'
DUPLEX
IMBRICATE
FAN
SOUTHWARD WITH "TRIANGLE ZONE"
DECREASING JP
ON THRUSTS 0°. .c.= I.E. TERMINATION OF
OR
44: PASSIVE ROOF DUPLEX WITH
ERODED
AND REACTIVATED :6? FORELAND SLOPING DUPLEX
(?) I eel
DUPLEX o OR ANTIFORMAL STACK
GEOMETRY
SHARP
(FAULT-
PROPA GATION
FOLDS ?) 0:4
NORTH ER N LIMB
Of SC)" R.
Fig. 12: Interpreted deformation styles in the NPDZ and beneath the
northern limb of the Soan syncline. See text for discussion and
figure 5 for location.
41

tions from either tilted strata or thrust faults. Between the opposed
reflections is a transparent zone (far left, fig. 11). This pattern
has been referred to as a "triangle zone" (Lillie et al., 1987); how-
ever, the genetic connotations of this term are ambiguous and in need
of clarification. The "triangle zone" is similar to that described by
Price (1981) and Jones (1982) for the foothills of the Canadian
Rockies, consisting of foreland-dipping reflections representing a
tilted layered sequence above an upper-level detachment, and hinter-
land-dipping reflections from blind thrust slices beneath the upper-
level detachment. A problem with this interpretation is that it
requires the presence of a south-dipping, upper-level detachment
surface. No south dipping thrust faults have been mapped in this
area. However, because the upper-level detachment could follow bedding
planes, it may be difficult to detect from surface geology (Jones,
1982). Seismic profiles (fig. 11) show that the lower part of the
Murrees thickens dramatically beneath the northern limb of the Soan
syncline, as compared to thicknesses penetrated by wells to both the
north and south. This thickening may be indirect evidence for the
upper-level detachment surface.
Seismic sections also reveal overlapping of concave-downward,
platform reflections near the southern margin of the NDPZ (fig. 10).
Drilling, which resulted in the discovery of the Dhurnal field (fig. 3)
confirms that these arcuate reflections represent horses of platform
rocks deformed into anticlines (figs. 8 and 13). Unfortunately, seis-
mic data gaps and transparency of the "triangle zone" prevent a com-
plete interpretation of the structural pattern. This pattern is
unlike the arrangement of horses in the northern part of the NDPZ, or
the stacking of thrust sheets at its southern margin, west of the
lateral ramp.
Analysis: Figure 12 depicts the preferred interpretation of deformation
style across the NPDZ. Thrust offsets of platform strata, observed on
reflection profiles in the northern half of the NPDZ, probably do not
die out into the folds described at the surface in the NPDZ. This

would require a rapid loss of displacement towards the surface, imply-


ing large scale ductile shortening in the hanging-wall rocks (Vann
et al., 1986). An upper-level detachment surface, above the hinter-
42

SOUTHERN POT WAR PLATEAU


"M BT" NPDZ SOAN SYN.

20 km (SOAN TROUGH, SALT RANGE, LILLA) + 45km (NPOZ) 65km TOTAL SHORTENING

200 km 150km 100km 50km 0 km

Fig. 13: Main (regional) balanced cross-sections and restorations. (A) Post-
Eocene/pre-molasse restoration (approx. 35-25 mya) or salt-wedge model.
Assumes flat top of Eocene. Note platform strata above jagged line are
interpreted to be eroded during subsequent passage of crust across the
peripheral bulge and associated basement faulting. (B) Post Middle
Siwalik/pre Upper Siwalik (approx. 4-5 mya) restoration. This restoration
depicts the foreland basin after basement faulting and prior to compres-
sional deformation. Deformation in the NPDZ and salt flow are not
pictured, but may have already been initiated. Basement dip determined by
assuming a 0.50 surface slope towards the depocenter (Baun syncline?) and
restoring the intervening strata. Formation contacts and faults are dashed
above the upward limit of preserved strata in (A) and (B). (C) Balanced
cross-section depicting present-day structure. Dashed lines above topo-
graphic surface represent eroded sediments. Eroded sediments and footwall
strata are not restored in the NPDZ and south of the Salt Range, respec-
tively. Dashed lines beneath topographic surface in the northernmost
Potwar Plateau indicates the questionable nature of the interpretation.
Anticline, just south of the NPDZ, interpreted in a gap in seismic coveage
(fig. 3). See figure A5 for a detailed interpretation of the Salt Range
portion of this cross-section. See figure 5 for location. Formation
symbols are the same as in figure 6.
43

land-dipping duplex is therefore suggested by the contrast in deforma-


tion style seen at the surface and observed, via seismic, at depth.
Cross-section balancing suggests that duplexing represents over 50%
shortening (fig. 13). Molasse above the upper-level detachment must
also be shortened by 50%. The deformation style of strata above the
upper-level detachment is unknown; however, these strata are often
eroded as they are elevated by underlying duplex thrusting (Banks and
Warburton, 1986). Regardless of the upper-level deformation style,
discrepancies between surface and subsurface structural style suggest
that strata above and below the upper-level detachment behave as
different lithotectonic units (e.g. Mitra, 1986).
In the southern half of the NPDZ, west of the oblique ramp, local
disharmonic folding (fig. 9) is explained by the stacking of hanging-
wall synclines above anticlines of the underlying thrust sheet during
imbrication, consistent with an imbricate fan interpretation (fig. 8).
Alternatively, this pattern is consistent with the preservation of
synclines above passive roof thrusts, suggesting continued duplexing
(e.g. Banks and Warburton, 1986). East of the ramp, the precise
structure within the "triangle zone" is difficult to interpret, because
the geometry of the structure disperses seismic energy. If the upper-

level detachment is a backthrust, and overlying strata have remained


relatively stationary, then this type of structure is the frontal end
of a passive roof duplex (Banks and Warburton, 1986; Vann et al.,
1986). In figure 13, the geometry of the "triangle zone" is constrain-
ed because thrust slices, exposed at the surface, are folded by under-
lying thrusts (see also fig. 8). Another constraint is that the roof
thrust slopes toward the foreland over a duplex height exceeding an
individual ramp height. In addition, the exposure of Eocene strata
suggests local triple stacking of horses (fig. 13). Therefore, the
geometry within the "triangle zone" consists of either partially
overlapping ramp anticlines indicating a foreland-sloping duplex
(Mitra, 1986) or completely overlapping ramp anticlines (stacked
horses), suggesting an antiformal stack (Boyer and Elliot, 1982). A

hinterland-dipping duplex with equally spaced faults is transformed to


these geometries by increased displacements on younger thrusts (see
central portion of NPDZ, fig. 13). In other words, instead of
44

propagating a new thrust, movement continues on the previous thrust as


would occur if the ease of sliding increases (Mitra and Boyer, 1986).
Therefore, a southward facies change toward weaker rocks (more salt)
in the decollement zone, may be responsible for the foreland-sloping
duplex or antiformal stack geometry.

Transition from the NPDZ to the Soan Syncline

Description: Along the southern boundary of the deformed zone is the


sharply upturned, northern limb of the Soan syncline. As seen on
seismic secttions (fig. 11) and supported by drilling, the near-surface
structure of this transition zone consists of a foreland dipping mono-
cline of molasse strata (fig. 13). This type of structure is common in
mountain fronts around the world (Vann et al., 1986).
Underlying the south-dipping molasse, platform rocks either con-
tinue subhorizontally (as in fig. 11), or are parallel to the structure
of the molasse. Where platform rocks are folded along with the
molasse, the transition from thrusting to nondeformation is marked by
folding. There is a gradual decrease in deformation intensity as
anticlines show reduced structural relief and the frontal thrust has
significantly less throw than thrusts to the north. In this case, the
frontal monocline (north limb of the Soan syncline) is the south limb
of an anticline cored by thickened salt (e.g. Khaur anticline, fig.
5). This pattern is only observed west of the lateral ramp.
Diverging platform and molasse reflections are observed on all
available lines east of the lateral ramp (fig. 11). Gaps in available
seismic coverage and ambiguous seismic interpretations do not exclude
the presence of this style of deformation west of the oblique ramp.
The continuity of platform reflections beneath folded molasse may be
explained by an intervening, upper-level detachment surface. The

presence of an upper-level detachment is consistent with the previous


interpretation of a "triangle zone", as discussed above.
Beneath platform strata, the transparent Eocambrian layer thins
eastward from about 0.9 to 0.3 seconds (i.e. about 2.0 to 0.7 km) over
a distance of 22 km; the molasse basin appears to thicken as the salt
thins eastwards. Evaporite thinning corresponds to the area above and
45

east of the interpreted position of the oblique ramp (fig. 6). Due to
lack of strike lines and poorly known velocities, it is not known if
salt thinning is associated with basement faulting.
Analysis: The interpreted oblique ramp does not appear to function as a
tear fault along which the northern Potwar frontal thrust west of the
ramp is displaced to the south; the branch line of the upper-level
detachment to the east is along strike with the Khaur anticline to the
west (fig.6). West of the oblique ramp, anticlines may be fault-propa-
gation folds (Suppe, 1985) in which folding precedes faulting and
thrust-shortening is replaced by folding upsection. East of the
oblique ramp, duplex thrusting appears to die out beneath a major
back-thrust or upper-level detachment. Strata above the upper-level
detachment are passively elevated by the southward driven wedge of
deformed rocks to form the forward dipping monocline.

Southern Potwar Plateau/Soan Trough

The Soan trough is a molasse basin containing a large thickness of


Lower and Middle Siwaliks occupying the southern Potwar Plateau. The

general shape of the Soan trough is that of a broad, asymetric syncli-


norium with a steep north limb and a gentle south limb bordering the
Salt Range uplift (fig. 13). In the area of study corresponding to the
central Potwar Plateau, the Soan trough is divided into the Soan and
the Baun synclines by an intervening anticlinal trend (Joya Mair and
Balkassar structures; fig. 14). To the west, where uninterrupted by
major anticlines, the Soan syncline encompasses the entire south half
of the Potwar Plateau and is synonymous with the Soan trough. In the
central Potwar Plateau, post Eocene sedimentary fill within the Soan
trough consists of a southward thinning wedge of Rawapindi molasse
(approx. 1.5 km in the Soan syncline and 0.8 km in the Baun syncline),
a large thickness of Lower and Middle Siwaliks (over 3 km in the Soan
syncline), plus smaller amounts of Upper Siwalik and post Siwalik
sediments, including the Potwar Silt.
Reflection data and surface mapping indicate that the Soan trough
is relatively undeformed. Structure within the trough consist of broad
folding with wavelengths of approximately 20-30 km (see Lillie et al.,
46

N BALKASSAR JOYA MAIR S


ANTICLINE ANTICLINE
A.
...-S

..
.rv
"I
- wl
nl

-w.

n Iry ...M+" .....wC^..M1w' =1T.wr.,1w.I!lM' -- ,y,wr


:, .. L

.- w. . r

~.11
T,i .. .. .w
`."yr.^w,u:w.w. nrd r
.
II ._Wr
« .b
.
L...

l I

5 km

B. r r w
,.;1 424
..-- ,- ..1,-
ti'iw nfi"sll
SM
r!
SM..a

w
W Ir
RM
-E -E
-E
SRF

PV
s::.'':
.nu":....'h_..w.u.'w' -. 111
!a
yam

II I
tJNN wL'
I...
L
_

,W,C nw'IU ..J I i ....,.


. SRF
r
Fig. 14: Uninterpreted (A) and interpreted (B), migrated seismic line
across the Balkassar /Joya Mair anticline complex in the south
central Potwar Plateau. This transect represents the only
discernable deformation in the south central Potwar Plateau. The
small syncline bisecting the broad anticline was possibly formed
by subsurface salt dissolution (e.g. Lohmann, 1972). Basement and
overlying reflectors are disharmonic, implying that supracrustal
rocks are detached from the basement. The decollement (dotted
line) is thought to be a diffuse zone involving salt flow, rather
than a discrete surface. Locally steep basement dip is perhaps
related to a basement fault interpreted north of the syncline.
Formation abbreviations are the same as in figure 9. See figure 3
for location. After Lillie and others (1987).
47

1987; their fig. 4). Folds in the southern Potwar Plateau consist of
a gently-folded, competent layer (platform strata and molasse), over-
lying a plastic layer of Eocambrian evaporites (fig. 14). Salt flows

from beneath synclines into the cores of adjacent anticlines, analogous


to folding beneath the Appalachian Plateau (Wiltshko and Chapple,
1984). Salt can have both an active and passive role in the formation
of folds. An active role implies that deformation results from the
buoyancy of salt (Martinez, 1974). Salt beneath the broad Soan trough
may have been squeezed southward towards the Salt Range due to increas-
ed isostatic pressure exerted by the molasse basin fill (Gee, 1983).
Salt may also have behaved ductily due to compressive stresses suggest-
ing a passive role (see Martinez, 1974).
Broad folding in the south-central Potwar Plateau may represent
the accommodation of slip at the rear of the thrust sheet (during base-
ment buttressing) by internal deformation. This process often occurs
in association with buried thrusts fronts, as would have been present
prior to frontal ramping. The Soan trough constitutes the trailing
portion of the Salt Range thrust sheet, which has been translated
southwards with very little internal deformation. This lack of defor-
mation is indicative of thrust surfaces that have propagated rapidly
and steadily towards the surface. Lack of intraplate and trailing edge
shortening represents an atypical style of development associated with
strongly emergent thrust fronts (e.g. Pine Mountain thrust; Morley,
1986) and is related to the decoupling properties of salt in the
decollement zone. The Soan trough appears to be analogous to the
Molasse Basin in the foreland of the Alps, which is a 50 km wide
zone of undeformed clastic deposits that have been translated 23 km
over Triassic evaporites (Mugnier and Vialon, 1986).
Deformation style throughout much of the western portion of the
southern Potwar Plateau is similar to the central area (Leathers,
1987), except for the zone of strike-slip faulting in the extreme west
(J. McDougall, in prep). However, the deformation style is quite
different in the eastern Potwar Plateau, where folds and thrusts are
aligned en echelon, in a SSW-NNE direction (fig. 5). Thrusts in the

cores of anticlines and exposed at the surface verge northward as well


as southward, while folding is characterized by broad, flat synclines
48

separated by narrower, more intensely deformed anticlines where


decollement splays have cut intermittently towards the surface (Johnson
et al., 1986). The seismic expression is remarkably similar to that
observed across the Parry Islands fold belt in arctic Canada, which is
also underlain by evaporites (Fox, 1983; Davis and Engelder, 1985).
Eastern Potwar Plateau folds appear to be fault-propagation folds
which are indicative of a slow rate of strain accumulation under low
stresses. Thickened salt in the cores of eastern Potwar Plateau anti-
clines is thought to represent layer parallel thickening, during
sticking of the propagating decollement surface (e.g. Fischer and
Coward, 1983). Therefore, the progressive build-up of strain along the
eastern edge of the thrust sheet is consistent with a slower rate of
thrust-tip propagation (Williams and Chapman, 1983) and may be explain-
ed by a stick-slip style of motion (e.g. Morley, 1986). Thrusts within
the cores of eastern Potwar Plateau folds may form because fold short-
ening is limited due to unavailability of salt for flow. In other
words, it is easier to thrust than keep moving increasingly viscous
flow zone material (due to facies change?). If there is an eastward
facies change to a less salt-rich composition, then the broader wave-
length folding of the central Potwar is explained by a greater visco-
sity contrast between the Eocambrian layer and overlying competent
strata (see Wiltschko and Chapple, 1984)
Flat topography, in combination with shallow basement dips beneath
the southern Potwar Plateau (avg. 2.70), indicates that the overthrust
wedge has a narrow cross-sectional taper. Narrow cross-sectional taper
and lack of internal deformation in an overthrust that extends far out
over the foreland are consistent with recent work by Davis and Engelder
(1985) on the mechanics of thrusting above a salt layer. For this
reason, the cause of contrasting deformation style and thrust-tip prop-
agation rates in the eastern and central Potwar Plateau is thought to
be a consequence of salt thinning eastwards resulting in an increase
in basal friction (Davis and Engelder, 1985). Alternatively, the
lessening of basement dip eastward (i.e. from ) 2.5° to (1°) requires
internal deformation to maintain critical taper (see Davis et al.,
1983) and may explain the change in deformation style (Jaume, 1986;
Jaume and Lillie, in press). A third possibility is that basement
49

ramps are responsible for the increase in internal deformation in the


eastern Potwar Plateau; an undulatory basement structure is interpreted
from gravity anomalies (Leathers, 1987), and reflection profiles
(E. Pennock, in prep.) across the eastern SR/PP.

Transition from the Southern Potwar Plateau to the Salt Range

Description: The northern flank of the Salt Range is an eroded mono-


cline; the south limb of the Baun syncline/Soan trough synclinorium.
Reflection profiles show that north-dipping sediments exposed at
the surface overlie a fault with the basement offset as much as one
kilometer, down-to-the north (fig. 15). Throw on the basement fault
diminishes to both the east and west, gradually dying out into basement
monoclines (figs. 5 and 16) The basement offset correlates with a
local steepening of the Bouguer gravity gradient (Farah et al., 1977),
and apparently acted as a buttress which controlled ramping and
emplacement of the Salt Range thrust sheet (Lillie and Yousuf, 1986).
Therefore, the monocline at the north flank of the Salt Range is the
surface expression of the footwall ramp. Similar ramps that are con-
trolled by basement offsets have been recognized beneath the forelands
of other collisional mountain belts (Lillie and Yousuf, 1986).
Seismic profiles and the detailed surface maps of Gee (1980) show
that the form of the northern flank of the Salt Range varies along
strike. Within the central area of study, the monocline can be divided
into three zones corresponding to different shapes. In the west-central
SR/PP, the molasse ramps gently to the surface at dips of 15-30° (fig.
17a and b). Farther to the west, surface dips are flatter, probably
due to ramping across a shallowly-dipping, basement monocline (Leath-
ers, 1987). The central regime begins east of the high-angle Kallar
Kahar fault (fig. 5), where the monocline is offset by a pair of
back-thrusts having a total throw of about 1.5 km (fig. 17c and d).
Farther east, the back-thrusts gradually die out as the monocline
steepens to 30° to 650.
Analysis: The north flank of the Salt Range was the location of com-
pressive flow and diapirism which are interpreted to have played
roles in pre-ramping deformation (fig. 18), as well as the development
Mb.

S
1,,,,...-.7 ....s... 1...1 ,:, t.Az i et......,-, ,,..7

,,4,
Al4 h.-."-- . ,..-.1.-.z1 'i.,;.r.-J-:'1, I- -ii ---,:t :.-f!...!- ti; le.:4.1.44
I 1%.,..-Itg,1-,.!,
'4,f,:44; t444,14:414.0: RA 0,41
r'lfr''' e-'1":"1%/lyli .4 4.'' :;;4'; ,e,71-ysloc.
li:=1,4..le..i....4.4;11. oN.:04, z.4:11-,,...;::A-r,-,1,;:,AYt. '2.-i .:011:1:;e7,-sti ..$:-*;','4*.e ;;;-14'' c-
is t,e..1: ,,,10'.??.., 1:,;14-.01? 4.:-..Q-7 44,-,-z---v- vps ill,
...t.,1: es.,441.1.%-t-A1 1-. .i..:44...t .411. mi-..14--:..:,. .i0.2-44,3ge.
..T4.44. 0.166.' - 1-----vAL ,-,-,-.1,-.9.:
::
:r i' c7;1 z:I.11qill''11 :411=='NF'4,'Pi*:;e1f.:14z44
4ki - --'-'4.41V. 'i, '..1q . I, 1'4% it . :I. -V"..0....!,1.e., - r,4,0; ." ----: ....:,. :4XIV
i.....ti.,4. 4,44 4t1 i: ti,tt-.S;
11/I V t I " . 0 .41 4/11;411.;.. 7../ls 1/ _.4 =-1;:j"... 11-- - --.:: i',. n, I Y :i1.1 cutt I
V_ ,ilk -^ SO l'.014 l'..-1.'i i.r ,-;* '-v. z 4;#,.--,.--- /' *.i -;4N
Et il 144"....,'"';,,11.:47,.;t4111' C.T...:XNA- 1-..A'. 'till 4,. '..Ail'1.4'.wit Ile/AWL" --41iFS1V-?' '41:.i: kl'i**1;441:-'1.rifr'it::
IF '1+.:' dp ;c.v., -4r i:c-'. -1=-:u:41.0 ..4' ir' tem );14^,71.; iv,;Nimie . . . .1.ZWto,
Airs 'f'il '......1'4'' 74 i 41.444'A, 1,-:.-444:;' t'ir' ii):0. 11.11'1.400; 1)13041glili g-41:11 'le e"j_.-.
`4*-1:.,,.. ' ir.1);77_1' z-4-41'41.44i 't; --illt".11',1,4 I 0, ,_i.P-1.,,,,;:: rorpowt .,..:,.... :,...-,:',;,;. t._1.....?,
7.!:.1.+1 %.,;,;:rzITI',e.:Ze-ii:7;,,;.:11.;; -.. ...11.'..:,- --'1' --:1' ' , L'''',L --f.----12".
1;.....t1T-:-...,..;Iffiti..,...,',.'....**1 4 IA - ,. -......- .,k, v.,..1:31,. !,:`..:!4I'
-4....-..... ;pi:14344 ,..... -,,jr. -.4- .4.., ............,...,f; , -...., . to:, 1.41;:i .1a,
51 /ri)11I .1414111. ,1111,1i4P V14'i,s?,;.';`,"zs,',g4sIlifil:kil.l'ititi'044- : 4 411,1?11: 4i.i'S:40.0:
411114' !IA 1 '1141-.--',,;.3.; 4:4; A ,;1g:-V,1,1;;;Ats.14.: , 044.00 kki4v,g4411, ?);.:1. Pr.., 9.1ic..-t
li 0 111,. '.4- 3 ,,,,,,r.; ,,,--;44-..:- , 1`,..q. : .,.- v 4 . INglitotp. $.. Oareom (414 'ATI': 2 -11rWi
-.11( 1 '11° , ilq,Z? .A;''. +i:,:c Iiilt,A., ',;A,p,'$,. ;T:44 Mt It T. ixootv $41.:.....i.44. 41.41e,
.16 ',,,cajj;.7i, t,(441;'N. f, 'fr,',41.114 0.?Irort, '0.".".474lkii; ilitImpiq ,;',--,.. 1..,1;`{4.0:-.r fit4,47.?
4.4,f:1- .4144JA 141,U,QVI. 4A::,,, ,41,..,21.4741.1-,-.11.4:yo.,..VA:;11,11;, 4s$9 .14 11.44 \ 440401,14.14.44,1i:$ii ',.--; #.4
KR ;g1:004.111 v+a-'. ' =IS A:14:11% tw,a it,ii a 2 ......ttta. ,..t.,.41v14:1g A41'11441/. 4...,-....,...44 ,,114444.0.4141
,. 1.11; ?,,k,,,1 4,161/40
1;41 411,14)VAT /11414$0:0 11141.:11;1}; is:7,1/N 4.111,.......11,,r_,104.1?,,y,.iy
':,,,z,vie It, li_,Z, ...21' I :.4..ks,, 11.3.33.1; .1, 11 ,t.,t- .. 1 1.._10,..,-.... )0 a
''/<`3 - h 011P+Tik:11".# -''.)2.

. II it: 'I, I II I I II I '

AJ H ;

t. , 11

') 1 4;1? 111/i '..:4 :Ifilit' ;


1 j iji i 1 j d '4 Ititli !:7%;-i:'
'-;404 IAN: liA 1 q11,1,di 44,-L9;."1::.4 ,1;:;:::")41c`:5:;" 34 ..):(t-:._:.;411,:....;
1.?0, 44t1,,,..,10,1:4,
.A0441; ,at 1i. )4-4.01.fiti 1": ',igiR4
gr.4.4, ,i'. I 7 t':.:.it$ ii , 4 .4 1Y,
if. 5. ,7_,,,,,,=-1 $9._ "73/4 . 1 . .
3.. .1.4 4:,..' 1;-.
"1":; ,-,,.

r;4:00"4151fuir'tr._-:---..-.-- 1.: : ...,-,--,-,.; e.,-,,;74.-",-...---,1_11-.C.'


'=4 -49 FA'. 41..;4P.4:4.4. . 4414.iet
1::.111;i7:1- ,E-..:'''' -.7': 04
46Y- J 1tF1
;0' 1.4J4110F°44.44,5 114111414fiii:fitAii
4, ,,l, 4
.14kai;;44'vlr'
5117, .-j-i -1;1:14 '23:144A:t itjj:14X
:;;1-404 , *.17,7444 )4.4!1/4 . 4 .4e.;CAST..
1 01 el.grilt.A
s 441V. .,,...4.4_,* Al t4e.44,10"; totomi<-1.- - 4Ittr,: falliIIP
r
01 MO 144:41.1.#444
-, 067::leit

'
,

, .........- -,444, ....A*


...I:of
...,4j 7.-grenfffel. Ja" .
..:411-44..' roe10,-,.71-P.,.......II/,-:iiiii."..""f":;41.11:11:1kr.:;:i.715441:12;vtr-th:

..
:11-1,,, 474: :1, .7.;04.4 42::::;2 411..": a7t..t ,;.,i14;f:II:j.1:;:1111itliii;;;41,11 i14-r;41:10 1;;;Ili 06:Ili 11;111,4;:i:"!1),:,:11.1.,;
.44',.'

In-114' 4.,.11:CIrlitS-1.77/N41'4441)
1 i-4
f? ( 4,14 I 4 g0041,4 V Y44 .41, 444411 4`i44f ' I'd 9.' .4 -4,
114,4 4 . 1 4 4
4
.,
::411111111P1111, 111$1114; SIn714i 0414;
:14142;e110 Y4itirerAig711.:itt itl,pl.,..' 4 ),I111.t. 1.A;;I:Sli,,,,,.-11,,,;,,
'WI-Iva
Al 1.1.r.oWit.-.';
,-2, IIIit.1:1012
.=1% 't-'iP. 0Pv,
fl. 1.:11Trt
--Ist.vp..41.
41.t4ga? 4 ,4 teigl 14 11)11,1012`
3-s.... ..... u r .. , 1 4.''''.! 14.-
-4144(ja'41,tie. 4,,. v -1
' , ' ' '''1 .4.-,,r.d, 1'. 14-44Ii:1?i r.:4,kviimon'.4i}:lk"411'411:37411.1'.

I .
I II 5

I II " Ill

- - 5 . II
:
. I II '.I - . "
a. : . . . "

III .. . ' . - I"


. I I . -I
51

Fig. 16: Geologic map of central Salt Range (after Cotter, 1933; Gee,
1980), showing lines of cross-section balancing (figs. 13 and
17). Note the westward increase in separation between footwall
cutoff (i.e. basement offset, portrayed by dotted line) and
preserved portion of hanging-wall ramp (range front); see text for
explanation. Amounts of shortening determined from section
balancing: A-A'=23.2 km; B-B'=21.2 km; C-C1=21.8 km; D-D'=21.9 km;
E-E'=23.2 km; and F-F1=19.7 km. Shortening values are less in the
reentrant between B and D, where erosion of the range front may
have been greater (see fig. appendix C). The lowest amount of
shortening (i.e. F-F') is interpreted to be in conjunction with
the erosion of imbricate thrust slices from the range front (see
dashed thrust faults). The southernmost dashed thrust depicts a
possible restored position of the hanging-wall ramp. The actual
position of the buried range front thrust is thought to lie
between surface exposures of allochthonous strata and the restored
position of the hangingwall ramp. Formation
abbreviations: Tp=Paleocene; and P=Permian; other abbreviations
are the same as in figure 4. See figure 5 for location.
7 2°30' 72°45' 73°00'
32°55' 32°55'

QAL
I
* 0%
0 0 40
00 US
co

DP, Na, Ch

RM
3 2°45'
= Tp -E

-C-- P

SRF

32 °33'
Figure 16
0 10km 20km 30km
NNW
1 1 1
SSE
TOPO. _-
Ch -1- 2km
SURFACE
_-? RM JA- -
LE IWO
SL
MS
SRF T -QM

--2km
SRF -C E
SRF

. MIM.0 --4km

Fig. 17: Frontal balanced cross-sections. Restored sediments dashed above topo-
graphic surface. Middle Siwaliks are not restored, due in part to uncer-
tainty in thickness change towards restored position of basement fault.
Note eastward (i.e. from A to E) steepening of hangingwall ramp across
basement offset and increase in internal deformation near the toe of the
thrust (especially south of KDF on E-E'). Dashed thrust splay on (E) de-
notes an alternative interpretation to high angle, strike-slip fault mapped
by Yeats and others (1987). Surface formation contacts, strikes, dips (not
illustrated) and faults are all from 1:50,000 maps of Gee (1980). See
figure 5 for fault and figure 4 for formation abbreviations and figure 16
for transect locations.
0 10km 20km 30km
NNW SSE

2km
TOPO
/SURFACE
SL

0 10 km 20km 30km
NNW KKF SSE

--- Ch 2km
ft- _7 7
firsi TOPO.
SURFACE
SL

2km

4km

Figure 17 (cont.)
30km D.

2km
TOPO -
SURFACE
SL

2km

4km

0 10km 20km 30km

2km

SL

-2km

-4km

Figure 17 (cont.)
56

Fig. 18: Ramping diagram demonstrating a possible mechanism of incorp-


oration of footwall strata as the ramp shallows. Horses/slices
of molasse (exaggerated in size to demonstrate the mechanism) are
plucked from the footwall and eroded near the toe of the thrust
sheet. Note (jagged) erosional surface through time. See figure
A5 for the initial and final configurations of strata across the
ramp. (A) Preramping deformation. Basement buttressing causes
buckling of competent strata above an elongate salt pillow or dome
formed in response to southward directed salt flow. Reactivation
of the normal fault plane, in strata overlying the basement,
resulting in reverse fault separation allows folding without foot-
wall deformation prior to ramping. (B) Incorporation and erosion
of the first molasse slice (south dipping stipples) and position
of next molasse slice (north dipping stipples). Dead zone (DZ)
grows as the ramp shallows. (C) Continued erosion and footwall
incorporation at the toe of the thrust sheet. Smaller slices of
Siwaliks and fanglomerates have been identified at the range front
(Yeats et al., 1984; Gee, 1980).
S

_1,, (,:ja.:-10. -",;!,;(2 -


. 'tit :-\ I
/- - I / Th
_ 1..
/
f.itOtt"-; ri 1\;:
,,
I I

,/-s. -
.
1-/.
01

11!:Z;-i I1 Z.11'.../%-, "" ""s \'- "


`T';,1>'-',"-,1-1,-,s'es,r,s1-',`
-

Figure 18 Ln
58

of the present monocline shapes. Pre-ramping deformation helps to


explain the emplacement of thick salt beneath the Salt Range and avoid
an unreasonably sharp initial interlimb ramp angle during subsequent
ramping. The basement fault, acting as a buttress to both the south-
ward salt flow (Gee, 1983) and the southward propagating decollement,
may have localized the growth of an elongate salt pillow or dome as the
overlying competent strata buckled against it (fig. 18). Regional
horizontal compressive forces cause evaporites to become mobile,
resulting in diapirism (fig. 18a; Martinez, 1974). As salt growth
continued, Cambrian and younger strata on the north side of the fault
would be lifted into contact with less competent molasse. Compressive
stresses could then push the thrust sheet across the buttress (fig.
18b). Once ramping was initiated, differential overburden would
cause salt to flow from the region of higher to lower load stresses
(Jenyon, 1985); i.e. southward from the downthrown fault block to the
base of the overthrust.
As the central portion of the thrust sheet between high-angle
faults pushed forward, the thick accumulation of salt would be accom-
modated beneath the anticlinal structure south of the ramp and by the
steepening of dips north of the basement offset (fig. 19). Reverse
faults probably didn't form until normal stresses built up as the
thrust lengthened (fig. 19b; Wiltschko, 1979). Reverse faulting is a
mechanism by which thrust sheets migrate through sharp, lower hinges
during ramping (Serra, 1977 mode 3; Moore, 1977), and apparently
occurred in conjuction with ramp steepening during diapiric salt
growth. Therefore, different footwall ramp shapes formed as a result
of salt accommodation and segmentation by strike-slip faults.

Salt Range

General Description: The Salt Range is a hanging-wall, ramp anticline


emplaced across a low-angle step in the sub-thrust surface. The gener-
al shape of the Salt Range is similar to Suppe's (1983) model of a
fault-bend fold, but without the frontal limb. Also, Suppe's precise
geometric relationships are not strictly applicable because of the
plastic behavior of salt at the base of the thrust sheet (fig. 19).
59

Fig. 19: Schematic diagram (to scale) illustrating the evolution of the central
Salt Range. Roman numerals represent progressive stages of ramping.
I.=early, II.=intermediate, and III.=late stage of ramping. Letters cor-
respond to: (A) east-central; (B) central; and (C) west-central slices
of Salt Range. Double barbed arrows indicate the interpreted
direction of
salt flow. Dashed line is the surface trace of the
original normal fault
plane. The footwall ramp shallows as ramping progresses. A reverse fault
develops to accommodate movement over the highest portion of the basement
butress in (B). In (C) footwall clipping allows shallow-ramping without
reverse faulting or a steep monoclinal flexure (as in A), and results in a
far-traveled horse. There is an increased amount of internal
deformation
from west (C) to east (A). Note that a large thickness of molasse has to
be unroofed before platform strata is subjected to erosion.
60

Alternatively, the Salt Range resembles a thrust nappe (e.g. Ramsay et


al., 1983; Charlesworth and Kilby, 1981); again, the frontal limb is
missing or thrust over during ramping. More accurately, the Salt
Range's anticlinal shape is the result of buckle or diapiric folding
prior to ramping and bending moment folding during ramping.
To the west, Salt Range structure becomes more complex, especially
near the western terminus of the range, where deformation is by a
combination of thrusting and right lateral, strike-slip faulting
(J. McDougall, in prep). The overall ramp anticline shape of the Salt
Range does appear to continue, bending to the north where its genesis
may be due to oblique or lateral ramping. The presence of more thrust
faults within the western Salt Range (see Gee, 1980) suggests increased
imbrication.
Beneath the central Salt Range, reflection profiles reveal an
autochthon of highly-reflective, probably Cambrian, and perhaps
younger, platform strata (fig. 15). Autochthonous molasse may also

exist beneath the Salt Range; this possibility is difficult to inves-


tigate because the seismic expression above the highly-reflective
sequence is chaotic to transparent. The amount of molasse at depth
depends on the orientation of the thrust surface, which is difficult to
discern because indisputable thrust surface reflections are absent
beneath the Salt Range. Locally, the Dhariala well penetrated over two
kilometers of Salt Range Formation (fig. 3; Gee, 1983; Khan et al.,

1986). This thickness of salt is probably anomalous because the well


was drilled in the zone of steepest ramping and upon a truncated
anticline. Thick salt suggests that the detachment remains at depth
and did not cut abruptly to the surface just south of the basement
offset. The decollement is not thought to have completely flattened
upon autochthonous platform strata, because a secondary mechanism would
be required to cause the thrust to ramp to the surface. Accordingly, a

moderately-dipping thrust is interpreted, resulting in a northward-


thickening wedge of salt and southward-thickening wedge of molasse
(figs. 13 and 17).
Although overthrusting with little internal deformation is common
in the west-central Salt Range, internal deformation increases to both
the east and west. To the east, folding is prevalent between the
61

Kallar Kahar strike-slip fault and KDF (fig. 5). South of the KDF,
there is intense deformation at the toe of the Salt Range thrust (fig.
17e), where nonparallel, locally overturned folds and reverse faults
are present (fig. 17e).
Although the frontal thrust is buried by fan gravels and alluvium
throughout most of the central Salt Range (fig. 16), range front
thrusting is exposed in the eastern and westernmost Salt Range (Yeats
et al., 1984). Folds and reverse faults within fan gravels, and
thrust faults between fan gravels, Siwaliks and older strata are
observed. The eroded fault scarp at the range front may actually be
the preserved hanging-wall ramp. Its scalloped outline is due to both
curved reverse faults and the concentration of erosion and alluvial fan
deposition at reentrants (Yeats et al., 1984).
South of Salt Range thrusting, sediments slope gently northwards
from the Sargodha High. Compressional deformation is present 15 km
south of the central Salt Range at the Lilla anticline, which appears
to be cored by a blind thrust within Eocambrian evaporites (fig. 13;
Lillie et al., 1987). This suggests that the deformation front may be
shifting southward from the Salt Range. Alternatively, the Lilla
structure may represent weak crumpling of footwall strata formed during
rapid emplacement of the Salt Range thrust sheet (e.g. Morley, 1986).
Eastward Termination of the Salt Range Thrust Sheet: The character of
the Salt Range changes southeast of the KDF (fig. 5). All available
reflection profiles crossing this fault and the area to the southeast
show thrusting and duplication of the platform sequence (see fig. 20).
Both forward and hindward verging thrusts are mapped, dividing the
eastern Salt Range into several fault blocks. Seismic lines constrain-
ing the position of autochthonous platform strata, indicates that
shortening is no longer concentrated on the range-front thrust (fig.
20). Complications are indicated by the mapped changes in vergence of
the KDF (fig. 5; Gee, 1980), which has been described as "flipping like
a propeller blade" (Voskresenskiy, 1979). Farther east, where the

range bends to the north, thrusting is replaced by active folding.


The KDF cannot be the eastward continuation of the Salt Range
frontal thrust, since it would require an abrupt reduction in shorten-
ing along strike (fig. 20). Rather, it may be a splay off of the
62

Fig. 20: Hanging-wall/footwall cutoff separation map of the east-cen-


tral Salt Range and Karangal/Diljabba (KDF) zone. Patterns denote
individual thrust fault surfaces between cutoffs. Patterned areas
represent duplication of platform strata, and therefore show the
minimum amount of horizontal shortening. Hanging-wall and foot-
wall cutoffs and duplication constrained by seismic lines; cutoffs
dashed where extrapolated beyond seismic control (fig. 3). An
eastward decrease in overthrust distance (and therefore shorten-
ing) on the frontal thrust is demonstrated. However, duplication
related to frontal thrusting is unknown in easternmost portions of
map (see large question mark). There does not appear to be an
eastward decrease in overall shortening. The KDF is the surface
expression of thrusting that accounts for a large amount of short-
ening. The change in polarity of the KDF is explained by overlap-
ping thrust sheets; eastern thrust sheets are interpreted to over-
lap western ones. Mapped positions of cutoffs may not coincide
with polarity flips because mapped cutoffs mark the base of plat-
form strata while younger strata are normally exposed at the
surface (see fig. A2). Northernmost cutoffs from Pennock (in
prep.). Minor thrust fault surface (less than 3 km overlap) in
north-central portion of map not shown (see E. Pennock [in prep.]
for detailed interpretation). See figure 5 for location and
abbreviations.
73°15'

6.04r0 THRUST MAPPED AT


SURFACE (GEE,I980)
HANGWALL CUTOFF
BASE PLATFORM STRATA
___,...../ FOOTWALL CUTOFF
BASE PLATFORM STRATA

1-1
\\
SALT RANGE FRONTAL THRUST

SE FORWARD VERGING THRUST

BACKTHRUST
1111111

NW FORWARD VERGING THRUST

3 2°45'

73° 73°15'
Figure 20
64

frontal thrust dominated by strike-slip offsets (see Yeats et al.,


in prep.) near the range front (fig. 17e). Figure 20 demonstrates an
eastward increase in the amount of shortening along the KDF zone.
Overlapping thrust sheets explain polarity changes along the KDF,
suggesting that it is the composite trace of several faults (fig. 20).
A back thrust (fig. 5, where teeth are on the southeast side of "fault
A") is seen to cut an earlier formed thrust to produce a "pop up"
structure, or "triangle zone" (as defined by Butler, 1982), indicative
of a slow rate of strain accumulation. By this interpretation the
transition from the central to eastern Salt Range may be analogous to
the transfer zone between en echelon thrusts described by Dahlstrom
(1977), although complicated by additional structures. Because both
thrusting, in combination withfolding, and strike-slip faulting are
interpreted, the KDF appears to be a zone of transpression.
Structures within the Central Salt Range: The Salt Range consists of
strong folding, faulting and uplift in a narrow zone. In the central
Salt Range most structures can be classified into five categories:
(1) minor folds with axes approximately parallel to the range front;
(2) high angle faults; (3) minor thrusts within the allochthon;
(4) diapiric salt structures; and (5) strike-slip fault zones.
The east-west orientation of minor folds (fig. 16) suggest that
they were formed by the same compressive stresses that emplaced the
Salt Range thrust sheet. Most of this folding appears to be concen-
tric, although nonparallel folding occurs locally, near the range
front (fig. 17e). The intensity of folding increases southwards,
towards the toe of the thrust, as well as toward the eastern Salt Range
(fig. 17).
High-angle faults within the Salt Range (fig. 16) are mapped as
"normal, or probably normal" by Gee (1980), although many appear to be
reverse faults formed by compressive stresses (fig. 17). Bending-

moment faulting is probably responsible for many of the high angle


faults. As a thrust sheet ramps up-section, its upper surface travels
through a compressive regime at the foot of the ramp, and then an
extensional regime where the thrust sheet flattens; the opposite is
true for its bottom surface (Wiltshko, 1979). Thus, brittle platform
rocks may experience normal faulting both at the foot of the ramp where
65

they are below the neutral surface, and again at the crest of the Salt
Range because platform rocks are above the neutral surface after over-
lying molasse is eroded (fig. 19). Normal faults may also be due to
extension above salt diapirs.
Other than large scale thrusting at the range front and along the
KDF zone, low-angle reverse faults within the Salt Range allochthon are
rare. These small scale thrusts (less than 5 km long), concentrated
near the range front (fig. 16), may be horses clipped from the foot-
wall, but most represent imbrication within hanging-wall strata. Sur-
face maps show thrusts and higher-angle reverse faults associated with
salt diapirism, both in gorges at the range front and along strike-slip
faults. Salt bursts up into planes of weakness where geostatic pres-
sure within the salt is greater than hydrostatic pressure in overlying
sediments (Ala, 1974). Inspection of geologic maps along the range
front (fig. 16) reveal transverse (i.e. north-south trending) folding,
consisting of sharp-crested anticlines in valleys, and flat synclines
making up topographic highs. This inverted topography is due to
gravitative flow processes (salt diapirism) related to unloading as
valleys are eroded by streams (Gardezi and Ashraf, 1974).
Three fault zones, which lie within the central Salt Range and
oblique to the frontal thrust have been mapped as active strike-slip
faults by Yeats and others (in prep). Landsat photos show about 3 km
of right-lateral offset along the Kallar Kahar Fault and 9 km of left-
lateral offset along the KDF. However, the offsets of individual
formation contacts are less on geologic maps (Gee, 1980), possibly
a result of fault drag. The three faults appear to function as tear
faults, segmenting the Salt Range; they separate different deformation
styles and cut structures not found on the opposite side (fig. 19).
Once the Salt Range thrust sheet was segmented, structures on either
side could develop independently (e.g. Dubey, 1980). Faults inter-
preted to have strike-slip motions are mapped as thrusts and high-angle
faults by Gee (1980). Within these fault zones, high-angle faults
bound horst blocks which locally bring salt to the surface (Yeats and
Lawrence, 1984). At places, strike-slip faulting is associated with
folding and reverse faulting, which collectively form flower structures
(fig. 17c). As previously noted, the KDF zone is dominated by
66

thrusting farther north (figs. 5 and 20); therefore, it may be part of


a transpressive zone. If so, then the central Salt Range is pushing
past its flanks and is a small scale example of "escape block tecton-
ics" as suggested by R. S. Yeats (personal comm., 1985).

Ramping of the Salt Range Thrust Sheet

Mechanism: Ramping and emplacement of the central Salt Range was con-
trolled by a down-to-the north basement offset acting as a buttress
(Lillie and Yousuf, 1986). For a discussion of other factors respon-
sible for the emplacement of the entire Salt Range, see Baker et al.
(in review, see appendix A). In the central Salt Range, a pre-existing
fracture through competent strata, if present, would have influenced
the ramping style. The presence or absence of this pre-existing
fracture depends on the timing of basement faulting (discussed below).
The proposed mode of ramping assumes post Eocene basement faulting
(fig. 13) and therefore a plane of weakness through platform strata.
Since the work of Rich (1934), hanging-wall folding has been
thought to be the result of strata moving over steps in the thrust
surface. More recently, it has been suggested that propagating
thrusts terminate laterally into and are preceeded by a "ductile bead"
of deformation (Elliot, 1976; Coward and Kim, 1981). For example,
the Pabbi Hills anticline (fig. 5) may be an eastward extension of the
Salt Range frontal thrust. Fischer and Coward (1983) reported field
evidence, in the Heilam thruat sheet of northwestern Scotland, indica-
ting layer parallel shortening resulting in asymmetric folding prior to
thrusting. Therefore, an alternative mechanism of hanging-wall ramp
anticline formation is possible in which thrusts cut through previously
folded strata, in this case formed by preramping salt growth, with
smooth trajectories similar to thrust nappe emplacement (Cooper and
Trayner, 1986). However, a lack of observable footwall deformation of
platform rocks (fig. 15) and the presence of a pre-existing ramp within
thrusted strata suggest a ramp-flat geometry (Cooper and Trayner,
1986). Noting Wiltschko and Eastman's (1983) work, it is likely that a
normal fault plane formed during basement faulting would be reactivated
in strata overlying the basement during compression and diapirism
67

resulting in a reverse fault separation (fig.l8a). If so, the southern


flank of the central Salt Range is not only an eroded thrust fault
scarp (Yeats et al., 1984), but also represents the preserved portion
of the downthrown normal fault block.
The shift from displacement along the reactivated normal fault
plane is an example of ramp shallowing. Serra (1977) reported field
evidence for ramp shallowing with increasing displacement. Steeper

ramps are abandoned in favor of shallower ones as weak rocks in the


lower ramp region of the upper plate thicken (see Serra, 1977, modes 1
and 2) as a result of salt flow (figs. 18 and 19). A shallow ramp
explains the absence of the frontal, forward dipping limb. Because the

Salt Range is young, it is unlikely that the absence of this limb is


due entirely to erosion. A frontal limb forms as a thrust flattens
after cutting up-section (Suppe, 1983). Therefore, a plausible explan-
ation is that the thrust doesn't flatten, but instead slopes gently to
the surface from the basement offset through relatively incompetent
molasse (figs. 13 and 17). Similar, shallow ramps are seen in the
foothills of the Canadian Rocky Mountains where thrust nappes cut
upsection through molasse (Charlesworch and Kilby, 1981).
Hanging-wall and footwall short cuts would modify the shape of the
ramp so that competent strata could move across the footwall with less
bending resistance (Knipe, 1985) producing a smoother and shallower
thrust trajectory (e.g. fig. 18c). Footwall clipping is an example of

a footwall shortcut (see fig. 17a). Wiltschko and Eastman (1983)


demonstrate that hanging-wall strata become immobilized or cut off
against basement faults during thrusting, forming a "dead zone"
(fig. 18).
Shallow ramping facilitated development of the great overthrust
length found in the central Salt Range. The resulting thrust sheet
profile is wedge shaped (fig. 17); wedges have a greater stable length
than rectangular shaped fault blocks (Murrel, 1981). At low dips there

is less work against gravity and bending resistance to continued thrust


sheet movement (Wiltshko, 1979). Work against gravity is further
reduced by the buoyancy of salt which helped to lift overlying strata
across the basement offset. The largest overthrusts are formed where
the basal rocks are the weakest, reducing drag (Murrel, 1981), and
68

where the toe is continually eroded, increasing the deviatoric stress


(Rubey and Hubbert, 1959; Rayleigh and Griggs, 1963). Therefore, the
presence of salt and unroofing of molasse probably enhanced thrusting
in the Salt Range.
Faulting of the basement in Eocambrian time would have created a
stratigraphic situation resulting in a different ramping mechanism.
Less southward salt flow would be required to emplace thick salt
beneath the Salt Range (fig. 21a). The basement offset could still
have localized salt growth due to convergence of flow over the basement
fault, because the salt would have thinned southward across the fault
(see Suppe, 1985). If pre-ramping deformation did occur, then the
structural style should be similar to fault-propagation folds of the
eastern SR/PP, implying footwall deformation. Because there would be
no surface of weakness for reactivation through competent strata, the
ramp would form by brittle fracture. In this case, a relatively low-
dipping ramp should result so as to reduce bending moment strain
and work (Wiltschko, 1979). Because a steep ramp is observed through
platform strata and no footwall deformation is seen (fig.15), this
ramping mechanism is not supported.
Removal of Molasse From Beneath the Salt Range: The interpreted shallow
footwall ramp through the molasse is due to two processes. First, the
thrust sheet climbs up-section as it overrides its own debris (fig.
19). The Salt Range thrust sheet is known to override fanglomerates
along the frontal thrust (Yeats et al., 1984). Although overriding of
fanglomerates is an important factor, it is not suggested that the
allochthon flattened onto a paleoland surface; restorations indicate
that Siwaliks may have been present on the upthrown basement fault
block prior to thrusting (see appendix B; fig. 13).
In addition, molasse may have been removed from beneath or incor-
porated into the Salt Range thrust sheet (fig. 18). Horses of Siwaliks
plucked from the footwall ramp during thrusting would be carried along
by the overlying thrust. Mountain fronts advance as footwall rocks are

incorporated (Elliot, 1976). Second order duplexing may be the method


of molasse incorporation; footwalls of major ramp anticlines are one of
the most common structural positions of duplexes (Mitra, 1986). Most
of the evidence of duplexing would have been removed by erosion during
69

Fig. 21: Restorations illustrating stratigraphic thickness changes


versus the age of basement faulting. Formation patterns are the
same as in figure 6 except Lower to Middle Siwaliks are subdivided
(upper Middle Siwaliks are cross-hatched). (A) Post Dhok Pathan
restoration depicting Eocambrian basement normal faulting. Salt
thickening (tabular salt model) by 1 km north of basement offset
supplies approximate area balance of salt. Basement must remain
approximately unflexed for forward topographic slope. Eocambrian
basement faulting does not explain local southward thinning of
platform strata. (B) Post Rawalpindi restoration showing early
molasse and platform strata thinning across the basement offset.
However, the Rawalpindi Group thins regionally to the south; the
estimated pinch-out is near the basement fault. Early to middle
Miocene basement faulting does explain the southward thinning of
platform strata. However, platform strata thinning can not
account for a kilometer of stratigraphic thickness change by
itself. (C) Post Dhok Pathan restorations showing the thinning of
(1) Lower to Middle (about 13-8 mya) and (2) upper Middle (about
8-5 mya; vertical-ruled) Siwalik molasse across the normal fault.
These two restorations correspond to basement faulting related to
Neogene flexure. Thinning of the upper Middle Siwalik molasse
across the basement offset (model C2) is most consistent with
restored stratigraphic thicknesses (fig. 13) based on estimated
positions of the basement offset versus the peripheral bulge
(e.g. fig. 7). However, it does not explain southward thinning of
platform strata unless platform strata was eroded along the crest
of the bulge prior to molasse progradation. An increased basement
dip on this restoration would result in a thinner Lower to Middle
Siwalik molasse section on the north side of the basement offset
(not shown). Southward thickness changes across the basement
normal fault could then be accomplished by a combination of
platform strata thinning (as in model B) and Siwalik molasse
thinning (models Cl and C2). Upper Siwalik and platform strata
thinning is consistent with the deposition of Upper Siwalik
conglomerate beds, containing recycled Permian Tobra Formation
clasts, interpreted as derived from the up-thrown normal fault
block (see fig. 22b). This latter version suggests basement
faulting at 4.5 mya (discussed below).
Iuum
71

ramping (fig. 18). Some horses of Siwaliks are preserved at the Salt
Range front. Overturned lenses of Siwaliks, identified beneath alloch-
thonous Salt Range Formation in the eastern Salt Range (Yeats et al.,
1984), are evidence of the incorporation of footwall strata by imbrica-
tion or duplexing. Footwall rock failure would occur in conjunction
with ramp shallowing (fig. 18), and is associated with fast displace-
ment rates (Knipe, 1985).
72

TIMING CONSTRAINTS ON EVOLUTION OF THE FOLD-THRUST BELT

Efforts to calibrate the development of the Himalaya are hampered


by the removal of sediments critical to dating structural movements. In
the foreland, a large effort has been put forth to date stratigraphic
horizons; tephrachronology and magnetostratigraphy have been success-
fully used to age calibrate units. This has opened the door to various
methods of constraining the timing of structural events (see Raynolds,
1980; Burbank and Raynolds, in press).

Migration of Folding and Thrusting into the Foreland

Although there is a general progression of deformation from north


to south across the modern foreland thrust belt in Pakistan (Johnson et
al., 1986), detailed studies reveal deviations from this classic
hinterland-to-foreland pattern (Burbank and Raynolds, in press). The

earliest documented deformation is pre-Paleocene displacement on faults


within the Attock-Cherat Range resulting in a low-angle unconformity
across the SR/PP (Yeats and Hussain, in press). A hiatus has been
interpreted to indicate that the Peshawar basin area was active during
the late Miocene and early Pliocene (Burbank and Tahrirkeli, 1985).
More recently, southeastward migration of the deformation front
has been documented along the eastern SR/PP (Johnson et al., 1979;
Burbank et al., 1986; Johnson et al., 1986). From northwest to south-
east, evidence for the southeastward younging of deformation consists
of (see fig. 5 for locations): (1) uplift associated with overthrusting
along the Riwat fault at the Kas Dovac site between 3.4 and 2.7 mya
(Raynolds, 1980; Johnson et al., 1982; Johnson et al., 1986); (2) Soan
syncline control of deposition at about 3 mya; (3) surface expression
of the Buttar fold after 5.5 mya; (4) surface relief of the Jabbar
anticline more recently than 3 mya (Raynolds, 1980); (5) unroofing of
Eocene strata at 2.5 mya along the Domeli Ridge (Johnson et al., 1982,
Johnson et al., 1986); (6) deposition of angular Siwalik clasts indi-
cating deformation of the en-echelon Lehri and Mehesian anticlines
at approximately 2.3 mya (Raynolds, 1980); (7) post 2.4 mya initiation
of deformation and post 0.7 mya surface expression of the Chambal Ridge
73

(Johnson et al., 1979); and (8) initiation of deformation along the


Pabbi Hills anticline at about 1.2 mya, with surface expression after
0.4 mya (Keller et al., 1977). Out-of-sequence deformation is exempli-
fied along the north limb of the Soan syncline, 5 km south of Rawalpin-
di (fig. 5), where tilted strata deposited 2.1 mya are overlain by flat
lying strata deposited about 1.9 Ma (Raynolds, 1980). These data pre-
cisely bracket an uplift event and, when compared to the ages of defor-
mation farther south, demonstrate a backstepping of strain release.
Yeats and others (1984) document deformation after 0.4 mya in
the eastern Salt Range. They suggest that thrusting along the Salt
Range front is active at a recurrence interval of thousands of years.
An active Salt Range frontal thrust is supported by micro-earthquake
studies which show that the entire Salt Range is active at low magni-
tude levels (Seeber and Jacob, 1977). The timing of initiation of
central Salt Range ramping is discussed in Baker et al. (in review; see
appendix A).
In the central Potwar Plateau, the Dhok Pathan Formation, dated as
5.1 m.y. at its upper boundary (Johnson et al., 1982), is folded and
truncated along the northern limb of the Soan syncline. Therefore,
NPDZ deformation is atleast as young as 5.1 m.y. There is no evidence
of active thrusting in the NPDZ. However, more interior portions of
the foreland fold-thrust belt may remain active, or are reactivated,
from time to time. This is exemplified by accelerated uplift of the
Attock Range after 0.6 mya and continuing tectonic deformation within
the Peshawar basin (Burbank and Tahirkheli, 1985).

Decollement Propagation Rate Calculation

The rate of deformation front migration or decollement tip propa-


gation is clouded by out-of-sequence deformation. This rate can be

estimated along the eastern margin of the SR/PP. Truncation of the


stratigraphic records at Kas Dovac and Pabbi Hills (fig. 5), indicating
the initiation of surface expression along the Riwat fault and Pabbi
Hills anticline, respectively, represent the most widely spaced equiva-

lent events. The events occur about 3 m.y. apart, in locations separ-
ated by 90 km, indicating a deformation front migration rate of 30
74

mm/yr. This rate is the same as the facies progradation rate obtained
by Raynolds and Johnson (1985) for the same area.
Due to sparce timing data and varying rates of decollement propa-
gation, an average migration rate cannot be obtained by similar calcu-
lations in the central study area. Instead, Williams and Chapman's
(1983) slip/propagation rate method is used. By dividing the derived
frontal shortening rate, which is equivalent to the slip rate, of 11
mm/yr (Baker et al., in review; see appendix A), by the percentage of
shortening, corresponding to the area over which the slip rate was
determined, then multiplying by 100, the propagation rate is obtained.
Cross-section balancing indicates approximately 20% shortening (e.g.
fig. 13) between the trailing branch line of the Salt Range thrust
sheet in the central Potwar Plateau (fig. 8) and the present deforma-
tion front at the Lilla structure; therefore, the resulting propagation
rate is 55 mm/yr. It appears that the decollement propagates into
undeformed foreland sediments at a rate temporarily exceeding plate
convergence rates, analogous to fault-plane slip rates exceeding plate
movements during earthquakes. This represents a fast rate of deforma-
tion front migration and is explained by decoupling along the salt-
lubricated decollement. By extrapolating this propagation rate back in
time from the present deformation front, continuing deformation in the
central portion of the NPDZ, until just before ramping at about 2 mya,
is implied. Care must be taken in interpreting these results because
the propagation rate is very sensitive to the derived shortening rate.
An important implication of the faster propagation rate is that defor-
mation would be occurring in the eastern Potwar Plateau, before termi-
nation of deformation in the central portion of the NPDZ.

Motion of the Thrust System

The southward migration of the deformation front across the


Pakistan foreland, in the central SR/PP portion of the fold-thrust
belt, appears to have occurred at a varying rate. Initial deformation
south of the Hill Ranges is assumed to have taken place in the NPDZ.
Intense deformation there is indicative of a relatively slow rate of
decollement propagation. At some point, the decollement began to
75

develop in the salt layer and propagated rapidly across the Soan
trough, before encountering the basement buttress. This added resis-
tance caused the sole thrust to stick and temporarily cease forward
propagation. Internal deformation of the thrust sheet, including
interpreted buckling against the basement buttress (fig. 18), took
place at this time. As the Salt Range thrust sheet overrode the
basement offset, slip was again transferred to the frontal thrust,
which ramped to the surface. In contrast, deformation style and timing
data (described above) suggests a more steady rate of deformation-front
migration in the eastern SR/PP.
During basement buttressing, slip at the rear of the thrust system
may have been accommodated by a continuation of thrusting or out-of-
sequence deformation in the NPDZ. While the deformation front surface
was continuing to step out into the foreland to the east, deformation
continued in the central portion of the NPDZ because strain was trans-
ferred northward from the buttress. Assuming equal amounts of shorten-
ing along strike south of the MBT, this interpretation suggests that
eastern Potwar Plateau shortening is not compensated by overthrusting
in the central Salt Range, but rather by NPDZ deformation. This

hypothesis is supported by the eastward narrowing of the NPDZ (fig. 5),


suggesting an eastward decrease in shortening there. Modeling by Mandl
and Shippam (1981) suggest that an increased push or buttressing will
cause fracturing, resulting in thrust ramping near the rear of a thrust
sheet. In the central Potwar Plateau this would be manifested by the
incorporation of additional horses or imbricate slices at the southern
margin of the NPDZ.

Tectonic Rotations

Early paleomagnetic measurements on Paleozoic rocks exposed in the


Salt Range were used to both constrain the overall rotation of Greater
India during collision and infer tectonic rotation of the Salt Range
(see Klootwijk, 1979). Both Paleozoic and Siwalik paleomagnetic
measurements suggest more than 6.5° counterclockwise rotation of the
entire Indian plate during the past 10 m.y. (Molnar and Tapponnier,
1975). From these same data over 200 counterclockwise rotation of the
76

Salt Range relative to the Indian shield has been suggested (Crawford,
1974; Klootwijk, 1979).
Opdyke and others (1982) divided new paleomagnetic observations
into two groups corresponding to declinations ranging from 3500-360°
and 3200-330°, respectively (see fig. 5 for locations). The first

group can be accounted for by rotations of the Indian plate. Sediments


from the second group have an average rotation of about 35° counter-
clockwise, well in excess of Indian plate rotation. By sampling
sediments ranging from 12 to 2 m.y. in age, Opdyke and others (1982)
found evidence suggesting that rotations were not contemporaneous
and concluded that the SR/PP is segmented into regions with different
times of rotations.
Burbank and others (in press) sampled along a transect crossing
the Soan syncline south of Rawalpindi (Soan and Kas Dovac sites; fig.
5). Sampled sediments range in age from 9 to less than 0.7 n.y.
Burbank interpreted a temporal variation in rotation rates and divided
the progessive rotation into three phases. During the first phase,
which covers an interval from 9-2.5 mya, 10-20° rotation took place,
representing Indian plate rotation plus regional rotation away from the
HKS. This is followed by little or no rotation during deformation of
the Soan syncline between 2.1 and 1.9 mya. The remaining 20-25° of
rotation took place rapidly after the deposition of the Lei conglomer-
ate (approx. 1.8 mya), and before deposition of the Potwar Silt.
Burbank claims that late-stage rotation is not due to local deformation
but rather a regional event above the decollement surface in response
to drag along the Jhelum Reentrant. Shortening values, as determined
by cross-section balancing, across the central Salt Range (fig. 18) do
not support a regional counterclockwise rotation of the Salt Range
thrust sheet during ramping, (discussed below).

Timing of Basement Faulting

In order to control ramping, the basement offset must have been


present prior to the encroachment of the thrust front. Possible
origins of the basement fault can be found in Baker et al. (in review;
see appendix A). The lack of sufficient stratigraphic thickness
77

changes and the presence of a substantial crustal thickness (Farah et


al., 1977), suggesting that the crust has not been thinned, seems to
rule out Mesozoic rifting as the primary cause of the present basement
offset. Similarily, basement faulting during the late Cretaceous
through Paleocene is unlikely, because contemporaneous strata are not
sufficiently thick to change thickness by a kilometer across the fault,
and preserved strata do not thicken towards the restored position of
the down-thrown block. The basement fault could have formed entirely
during the Eocambrian or the Neogene, as evidenced by stratigraphic
thickness changes constrained by seismic and exploration wells.
An Eocambrian event is supported by the thick allochthonous salt
of the Salt Range restoring to the down-thrown side of the basement
fault (fig. 21 a). Eocambrian faulting may have controlled the distri-
bution of salt. Fault control of salt deposition is also supported by
the change in deformation style in the eastern Potwar Plateau, attri-
buted to the thinning of the salt, coinciding with the dying out of the
normal fault. However, the presence of cratonic crust and lack of
seismic reflectors characteristic of synrift graben fill (Lillie and
Yousuf, 1986) do not support an Eocambrian event. Also, the preferred
ramping mechanism is inconsistent with Eocambrian basement faulting, as
discussed above.
The basement offset could also have formed entirely during the
Neogene, because the molasse thins by over a kilometer from the Baun
syncline to the Jhelum plain (Baker et al., in review; see Appendix
A). A variation of this theme would be late Paleogene to early Neogene
faulting represented by thinning of platform strata and Rawalpindi
molasse (fig. 21 b). However, Rawalpindi thinning is thought to repre-
sent a natural stratigraphic pinch-out and platform strata thinning is
better explained by later erosion related to southward migration of the
peripheral bulge (Baker et al., in review; see appendix A).
Neogene faulting is thought to be related to lithospheric flexure
as thrust sheets are emplaced from the north. As the bulge passes
through the crust, the upper crust is placed in an extensional regime
in which normal faulting occurs (Isaacs et al., 1968). For this

reason, and since seismic expression suggests a normal rather than a


reverse fault (fig. 15), reverse faulting related to the present
78

compressional regime is considered unlikely. Flexural modeling by


Duroy (1986; Duroy et al., in press) supports the interpretation of the
Sargodha high as a flexural bulge with an associated extensional regime
in the upper crust. Additional support of flexural extension of the
Indian continental crust comes from observations of two large normal
fault earthquakes which have occurred beneath the foreland of India
(Molnar et al., 1976; Ni and Barazangi, 1984).
A seismic profile across the Sargodha high, south of the Mianwali
Reentrant, supports its interpretation as a flexural bulge. Prominent
platform and basement reflections are warped into a broad, anticlinal
shape, indicating a post Permian and probably a post Eocene age of
development. Permian strata were penetrated beneath molasse by the
Kundian well (see Leathers, 1987). The northern limb is steeper,
consistent with the modeled shape of the flexural bulge (Duroy, 1986;
Duroy et al., in press). In addition, molasse reflections lap on to
the north limb as crust is flexed down into the foredeep. The sequence
of prominent platform reflections is truncated along the southern limb
and thinner on the northern limb, which is explained by erosion near
the crest of the bulge. Finally, horst and graben structures are
interpreted, from the reflection profile, along the crest of the
Sargodha high, suggesting normal faulting.
An Upper Siwalik conglomerate bed, deposited about 4.5 mya at the
Baun locality (fig. 16), contains clasts of Eocene carbonates and
recycled Precambrian clasts from the Permian Tobra Formation (Johnson
et al., 1986; Burbank and Raynolds, in press). These clasts are
important stratigraphic evidence constraining the timing of either
ramping or basement faulting. If the source area of Upper Siwalik
clasts is the Salt Range, then conglomerate deposition, in combination
with diminished sedimentation rates and differential rotation, may
indicate incipient Salt Range thrusting by 4.5 mya (Johnson et al.,
1986; Burbank and Raynolds, in press). Figure 22a shows that the
amount of uplift necessary to expose Permian strata as suggested by
Burbank and Raynolds (in press) would lift competent hanging-wall
strata (i.e. Cambrian to Eocene rocks) at the ramp above competent
footwall strata. In lieu of this, regional transmission of stresses
from the basement buttress resulting in the Riwat fault and MBT
79

Fig. 22: Schematic diagrams illustrating 2 possible modes of exposing


source beds of Eocene carbonate and Permian Tobra Formation clasts
found in Upper Siwalik conglomerates at Baun. The Baun site has
been "pulled back" 25 km northward, from the basement fault, to
account for later shortening on these 4.5 mya restorations.
(A) Allochthonous source beds uplifted by early activation of a
segment of the Salt Range thrust without overriding the older
basement fault (Burbank and Raynolds, in press). (B) Autochthon-
ous source beds exposed to erosion via uplift along a contempora-
neous basement fault. To expose beds immediately adjacent to the
fault requires an average basement dip of 5.5° or 8° (measured
from the down-thrown and upthrown basement blocks respectively);
these dips are anomalously steep compared to present basement
structure (see Duroy, 1986; Duroy et al., in press). Therefore, as
a slight perturbation of (B), autochthonous source beds may have
been exposed farther to the south along the crest of the peri-
pheral bulge at 4.5 mya, in line with more refined restorations.
An autochthonous source is supported by drilling at Lilla, where
source strata have apparently been removed by erosion (fig. 13).
80

movements (Burbank et al., 1986) makes little sense. In addition,

exposure of alLochthonous Salt Range source beds would require the


unroofing of approximately 2 km of molasse prior to 4.5 mya, while no
evidence of recycled Siwalik and Rawalpindi clasts have been reported
at Baun.
Alternatively, the erosion of clasts deposited at the Baun
locality are readily explained by passage of the flexural bulge. By

this interpretation, Eocene and Permian clasts are interpreted as


derived from the south, from the up-thrown basement fault block (fig.
22b). The Shell Lilla well (fig. 3) penetrated Cambrian rocks directly
beneath Siwaliks; erosion of intervening strata, including Eocene
carbonates and the lower Permian Tobra Formation, is therefore demon-
strated south of the basement offset. Baun conglomerate clasts are
interpreted as derived from an uplifted autochthonous source, directly
adjacent to the basement fault (fig. 22b) or located farther to the
south along the crest of the peripheral bulge (fig. 6b).
Differential rotations recorded at Baun are not explained by
passage of the peripheral bulge unless strike-slip movements accompan-
ied dip-slip faulting on basement faults, resulting in shearing of
overlying rocks. Strike-slip fault plane solutions of earthquakes
along the Sargodha High (Seeber et al., 1981) may be evidence of this
mechanism. Alternatively, rotation data or the interpretation of the
age of unrotated Upper Siwaliks could be in error. Magnetic polarity
stratigraphy depends on an unambiguous correlation between the magnetic
polarity time scale and the locally measured succession of reversals.
The correlation of the mostly reversed Upper Siwaliks at Baun is
admittedly ambiguous (Opdyke et al., 1979). For example the observed
succession could just as easily be assigned to the Gauss rather than
the Gilbert Chron (see Burbank and Raynolds, in press; their fig. 4),
indicating that Upper Siwaliks at Baun were deposited between 2.3 and
3.4 mya. Another possibility is that rotation occurred during early
initiation of thrusting in the western Salt Range (see Leathers, 1987),
which conflicts with Johnson and others (1986) interpreted time of
ramping.
If the Baun conglomerates are the last remnants of stratigraphic
thickening across the normal fault (fig. 22b), then basement faulting
81

was occurring at 4.5 mya. Otherwise earlier faulting is suggested


because it would explain a kilometer of Lower to Middle Siwalik molasse
thickening between the Baun syncline and the Jhelum plain (see fig.
21c). Using the derived rate of frontal shortening and the rate of
underthrusting beneath the Himalaya (see Molnar and Deng, 1984) as an
indication of the relative motions between Indian crust and the
flexural bulge, the past positions of the basement fault versus the
Sargodha high have been estimated (fig. 7). These reconstructions
suggest that the faulted basement was in the extensional regime at the
crest of the bulge, or along its sharply downflexed northern flank,
between 4 and 12 mya.
Another possibility is that an early formed basement fault has
been reactivated. A faulted upper crust could have been bevelled by
erosion prior to deposition of the Eocambrian and younger section and
then the fault reactivated during Neogene flexure. The great throw on
this basement fault as compared to other potential basement faults
identified on seismic profiles is thus explained by reactivation of a
previous plane of weakness in the upper crust.
The scenario of restorations constrained by salt retrodeformation,
paleobasement dips and topographic slopes is most consistent with
Neogene basement faulting (see appendix B). However, the controversy
over whether the basement normal fault formed in Eocambrian or Tertiary
time may not be easily resolved because: (a) Neogene thrusting altered
the original distribution of evaporites (Gee, 1983); i.e. salt thicken-
ing observed today across the normal fault could be due to salt flow
during compression and (b) erosion from the top of the thrust sheet as
it ramped to form the Salt Range has removed potential evidence of
abrupt molasse thickening and conglomerate deposition on the down-
thrown side of the normal fault.
82

SHORTENING AND ITS IMPLICATIONS

Amount of Shortening

Salt Range: Six balanced sections were constructed to determine the


minimum amount of frontal shortening across the basement ramp in the
central Salt Range and evaluate potential variations in shortening
along strike. Frontal shortening values, corrected to N 15° W, range
from 19.6 to 23.3 km (fig. 16). The greatest amounts of shortening
correspond to the sections farthest east and west. Therefore, the
balanced sections show little or no regional variation in shortening.
Because shortening doesn't vary substantially along strike and the
measured amounts of shortening represent minimum values, the larger
values are probably closer to the actual amount of shortening. Factors
which potentially effect the accuracy of shortening values are discuss-
ed in appendix C. The main factors thought to influence the derived
shortening values are erosional truncation and failure to account for
burial of the frontal thrust. Apparent shortening is less across
reentrants where erosion and deposition are greater; inverted topo-
graphy within these reentrants (see above; Gardezi and Ashraf, 1974)
suggests that present day valleys are areas of unroofing due to salt
diapirism. Therefore, erosion may explain differences in measured
shortening (fig. 16).
Southern Potwar Plateau/Soan Trough: The Soan trough is the relatively
undeformed, trailing edge of the Salt Range thrust sheet (fig. 13).
Only 0.4 km of shortening is measured across broad folds between the
axis of the Soan and Baun synclines, a distance of about 50 km. Short-
ening along the north limb of the Soan syncline is included with north-
ern Potwar shortening, while deformation of the south limb of the Baun
syncline is included with frontal shortening. Because of the decoup-
ling properties of the Eocambrian salt, Cambrian and younger strata in
the intervening area were translated about 23 km over basement with
almost no internal deformation. The 0.4 km of internal shortening
contrasts sharply with the area along strike in the eastern Potwar
Plateau where Johnson and others (1986) estimate "20% (i.e. about 20

km) shortening along the breadth of the detachment".


83

Northern Potwar Plateau: The calculation of shortening in the NPDZ was


hampered by poor surface control due to alluvium cover and lack of
detailed maps, lack of sufficiently deep drilling, and chaotic seismic
expression. The regional balanced cross-section suggests 45+15 km of
horizontal shortening of competent platform strata (fig. 13).

Frontal Shortening Rate and Comparison to Plate Convergence Res

Ramping of the central portion of the Salt Range thrust sheet


represents a discrete event from which timing and the amount of con-
traction of sedimentary cover rocks has been determined. Using the
amount of shortening from balanced sections and the estimated time of
ramping (Baker et al., in review; see appendix A), the minimum rate of
shortening between the Indian basement and the Phanerozoic sedimentary
cover is between 9 and 14 mm/yr.
Several factors have not been considered in this rate calculation.
Shortening within the trailing portion of the thrust sheet and across
the Lilla structure are ignored but add up to less than 0.5 km. In

addition, the reversal of an estimated 0.7 km of interpreted extension


during basement faulting (fig. 13), is also not considered. However,
these factors are offset by potential shortening prior to the thrust
sheet overriding the ramp (i.e. interpreted preramping deformation;
fig. 18).
For these reasons, and considering timing evidence, the earlier
end of the range of thrust ramp initiation is preferred (i.e. about 2.1
mya). Because cross-section balancing results in a minimum shortening
value and most sources of error would increase rather than decrease the
measured shortening (see appendix C), true shortening in the central
Salt Range is assumed to be at least 23 km. Therefore a frontal short-
ening rate of 11+2 km/m.y. is thought to be representative of the
amount of convergence absorbed south of the Hill Ranges (MBT) in
Pakistan (fig. 23). This rate is comparable to Lyon-Caen and Molnar's
(1985) estimate of 10-15 mm/yr of convergence being absorbed by under-
thrusting of the central Himalaya. The calculated frontal shortening
rate represents about 25% of the 42 mm/yr convergence between the
Indian and Asian plates at longitude 730 east (Minster and Jordan,
N
TOTAL = 42 mm/yr
I8-38mm/yr
STABLE
ASIAN
1--
Y HIMALAYA I-
1-4-11mm/yr-4.-
PAMIR f-
CRATON TIEN SHAN
rUPLIFT AND PB HR1' PP SR SH
PLAIN STRAIN ("ESCAPE BLOCK TECTONICS")
'INTERNAL DEFORMATION
AND UNDERTHRUSTING NORTH OF THE -7-27 mm/Yr SLIPri`rimm/yrZER 0 SLIP
HIMALAYA 4 24 mm/yr SLIP =18 mm/yt FRONTAL SHORTENING-4Imm/yr
SLIP=18 -38mm/yr
11'

MOHO
100 km

Fig. 23: Schematic crustal section demonstrating potential distribution of


convergence absorbtion. Minimum constaint on convergence absorbed north of
the Himalaya is 4 mm/yr (Molnar, 1987), while maximum is 38 mm/yr (Lyon-
Caen and Molnar, 1985). The other constraint is the derived rate of
frontal shortening, 11 mm/yr (Baker et al, 1987; see appendix A). The
balance of convergence (i.e. from a total of 42 mm/yr [Minster and Jordan,
1978]) is accommodated via internal deformation of the Himalaya. Slip
along the decollement is equal to the sum of convergence absorbed south of
the respective locality. Lyon-Caen and Molnar's 18 mm/yr rate of under-
thrusting is arbitrarily placed beneath the Main Mantle Thrust (MMT).
Crustal structure modified from Seeber and Armbruster (1979). HR=Hill
Ranges; PB=Peshawar Basin; PP=Potwar Plateau; and SR-Salt Range. See
figure 1 for other abbreviations.
85

1978). Therefore, most convergence is absorbed north of the frontal


thrust zone by contraction within the collision zone and escape-block
tectonics (fig. 23).

Tectonic Rotation Versus the Distribution of Shortening

Burbank and others (in press) mechanism of late-stage rotation


implies rigid counterclockwise rotation of the Salt Range thrust sheet
about an axis near the eastern edge of the salt basin. However, this
hypothesis is inconsistent with pre 4.5 mya rotation at Baun (Opdyke et
al., 1979; Burbank and Raynolds, in press; Johnson et al., 1986). To

test Burbank and others' hypothesis, the amount of shortening was


calculated for 20° to 30° counterclockwise rotation using various
moment arms extending to positions in the Jhelum reentrant; 29 to 94 km
shortening was required across the central Salt Range. The high value
is considerably more than measured by balanced cross-sections. The

lowest value, which is similar to estimates (fig. 16), was obtained for
a pivot near the eastern edge of the Salt Range. A pivot farther west
within the eastern Salt Range is consistent with shortening values.
However, a westward increase in shortening would be expected if
ramping were part of a regional rotation, as suggested by timing data.
No westward increase was found in the central Salt Range (fig. 16).
Also, the large amount of platform strata duplication and the potential
for additional shortening in the eastern Salt Range (fig. 20), suggests
that shortening there is close to or in excess of that measured in the
central Salt Range. Erosion should not have had a great effect on
shortening values because ramping occurred recently (fig. 19). An
increase in shortening in the western Salt Range to over 30 km, esti-
mated from a cross-section (Leathers, 1987) and supported by the
westward increase in the width of the premolasse outcrop belt, may be
consistent with gross rotation models. However, the amount of shorten-
ing across the western Salt Range is difficult to determine; the
location of footwall cutoffs, interpreted from reflection profiles, is
indiscernible.
Timing data can further constrain potential regional rotation.
Deformation at the Domeli and Lehri/Mehesian structures (fig. 5) is
86

interpreted to have preceeded ramping (Johnson et al., 1986). The

remaining amount of shortening, across the frontal thrust in the east-


ern Salt Range (fig. 20), appears to be significantly less than in the
central Salt Range. Add to this the higher estimates of shortening in
the western Salt Range, and a regional counterclockwise rotation during
ramping is implied. The timing of deformation along the KDF versus
frontal ramping and proposed regional rotation is important because it
represents a large amount of shortening. If ramping was part of a
regional rotation, then strike-slip faults segmenting the Salt Range
(fig. 5) may separate areas with different amounts of shortening.
Opdyke and others' (1982) interpretation of rotation is supported.
That is, rotation above the decollement is not rigid but segmented into
regions with different times of rotation. In this case, measured
rotations would be caused by local movements. Drag along the edge of
the salt basin can be used to explain rotation of eastern SR/PP
structures into their present allignment. These structures would be
rotated individually, in response to similar stresses, resulting in
their parallel orientations (fig. 5). In the central Salt Range,
different orientations of the Salt Range front and the basement
offset, indicate a westward increase in hanging-wall and footwall
ramp separation and are therefore evidence of rotation during ramping.
Instead of being the result of a regional rotation, the westward
increase in overthrust distance is locally compensated by an eastward
increase in internal deformation, mainly concentrated at the toe of the
thrust (fig. 17), suggesting a local rotation of the hanging-wall ramp
during thrusting.
87

CONCLUSIONS

Insight has been obtained as to the current structure and mode of


evolution of the central Salt Range and Potwar Plateau of Pakistan. A
pre-existing, down-to-the-north basement offset controlled ramping of
the encroaching thrust front. Palinspastic restorations, seismic
observations, and ramping style suggest that the basement offset is a
normal fault formed during Neogene flexure of the crust. Alternative-
ly, the basement offset may have been formed during Eocambrian rifting
and therefore controlled the original distribution of evaporites.
During later thrusting, the buttressing effect of the basement offset
caused a local buildup of salt. Lack of footwall deformation and the
presence of a steep ramp through platform strata support a ramping
mechanism in which the normal fault through competent platform strata
was reactivated in a reverse sense during initial stages of ramping.
Salt growth eventually lifted strata south of the offset into a posi-
tion where compressive stresses could push the thrust sheet along a
shallow ramp toward the surface. The thrust trajectory appears to have
been smoother than would occur during typical fault-bend folding.
However, bending moment stresses across the basement offset were still
responsible for passively folding competent strata into the present
anticlinal form of the Salt Range.
The thrust sheet north of the ramp in the southern Potwar Plateau
is only mildy deformed due to the decoupling properties of salt.
Farther to the north, the northern limb of the Soan syncline dips
steeply southward above an upper-level detachment. Strata above the
upper-level detachment and NPDZ thrust slices form a "triangle zone"
geometry. This style of deformation changes westward, across an
oblique ramp, to an apparent imbricate fan. The upper-level detachment
is interpreted to continue northward where it separates different
lithotectonic units in the NPDZ. Beneath the upper-level detachment,
competent platform strata are deformed into a hinterland-dipping
duplex.
The timing of deformation is poorly constrained throughout the
central Soan trough and NPDZ. However, changing sediment accumulation
rates used in conjunction with stratigraphic evidence suggest that the
88

basement ramp to the south was overridden about 2 mya. Evidence of


early to middle Pliocene deformation (i.e. Upper Siwalik conglomerate
clasts preserved in the Baun syncline), thought to indicate early
hanging-wall ramp uplift, is alternatively explained by basement normal
faulting, uplift, and erosion related to migration of the peripheral
bulge through the Indian lithospheric plate. Balanced cross-sections
suggest that a minimum of 19 to 23 km of shortening is present across
the frontal ramp, indicating a shortening rate of 9-14 mm/yr. There-
fore, frontal thrusting in the Pakistan foreland accounts for a
significant but not the majority of the total plate convergence.
Rapid decollement propagation along the salt lubricated decolle-
ment within the central Potwar Plateau was temporarily slowed by
basement buttressing. Differing thrust propagation rates between the
central and eastern Potwar Plateau suggest that deformation within the
central NPDZ was concurrent with shortening in the eastern Potwar
Plateau. Regional rotation of the thrust sheet may then have occurred
during subsequent frontal ramping as indicated by an apparent westward
increase in overthrust distance. However, central Salt Range sections
do not indicate a westward increase in shortening, as would be expected
as a result of proposed regional rotation of the Salt Range thrust
sheet. Instead, the westward increasing overthrust distance in the
central Salt Range appears to be locally compensated by an eastward
increase in internal deformation.
Previous foredeep and facies progradation rates appear to be over-
estimated due to the influence of recent rapid deformation front migra-
tion. Foredeep migration is instead controlled by the rate of under-
thrusting beneath the Himalaya, of which frontal shortening represents
the minimum rate, not total plate convergence. Palinspastic restora-
tions suggest that Lower and Middle Siwalik molasse deposition was
concentrated in the central Potwar Plateau, possibly by basin down-
dropping along the basement normal fault. As the deformation front
shifted to the Salt Range area, the depocenter jumped to the south
along the Jhelum River.
89

Future Work

Although many of the problems in the Pakastani foreland have been


addressed (e.g. Gee, 1983), future work is needed to fully understand
the structural development of the SR/PP. Additional timing information
would be helpful. For example, an indisputable age of the sediments at
Baun may help to further constrain the time of Salt Range thrust sheet
ramping and normal fault formation. The age of sediments at the Lilla
structure would determine its relationship to Salt Range thrust ramp-
ing. Similarily, the maximum age of the Potwar silt is needed to date
the initial uplift of the Salt Range. Additional rotation data would
help to resolve the controversy concerning rigid regional (Burbank, in
press) or segmented rotation (Opdyke et al., 1982).
Better surface maps, seismic data and additional well control are
needed to further resolve the structural style in the NPDZ. A branch
line map is essential to understanding the interconnections of faults
there. More accurate shortening data are needed in the NPDZ, and in
the eastern SR/PP to determine the regional variation of shortening and
its bearing on regional rotation. To this end, additional subsurface
data are needed along the northern flank of the western Salt Range to
locate footwall cutoffs and resolve the emerging controversy over the
amount of overthrusting preserved there.
In light of changing basement dips as a result of lithospheric
flexure, a dynamic model of the mechanics of foreland thrusting is
needed to study past movements of the thrust system. Similarily, a
revised model of foreland deposition should be developed that takes
changing basin shape and the effect of the Sargodha High into consider-
ation. Finally, the rate of deformation front migration plus the rate
and percentage of shortening should be studied in the Indian portion of
the Himalayan foreland fold-thrust belt where a salt lubricated
decollement is lacking. A comparison of results to the present study
could then be used to evaluate the decoupling effects of salt and the
origin of the HKS.
90

BIBLIOGRAPHY

Acharyya, S. K., and K. K. Ray, 1982, Hydrocarbon possibilities of


concealed Mesozoic Paleogene sediments below Himalayan nappes
reappraisal, Amer. Assoc. Petrol. Geol. Bull, v. 66, p. 57-70.

Ala, M. A., 1974, Salt diapirism in southern Iran: Am. Assoc. Petrol.
Geol. Bull., v. 58, p. 1758-1770.

Baker, R. A., H. M. Gehman, W. R. James, and D. A. White, 1984, Geolog-


ic field number and size assessments of oil and gas plays, Am.
Assoc. Petrol. Geol. Bull., v. 68, p. 426-432.

Baker, D. M., R. J. Lillie, R. S. Yeats, G. D. Johnson, M. Yousuf and


A. S. H. Zamin, in review, Development of the Himalayan frontal
thrust zone: Salt Range, Pakistan, Geology.

Banks, C. J., and J. Warburton, 1986, 'Passive-roof' duplex geometry in


the frontal structures of the Kirthar and Sulaiman mountain belts,
Pakistan, Jour. Struc. Geol., v. 8, p. 229-238.

Barndt, J., N. M. Johnson, G. D. Johnson, N. D. Opdyke, E. H. Lindsay,


D. Pilbeam, and R. A. K. Tahirkheli, 1978, The magnetic polarity
stratigraphy and age of the Siwalik Group near Dhok Pathan vil-
lage, Potwar Plateau, Pakistan, Earth Planet. Sci. Lett., v. 41,
p. 355-364.

Behrensmeyer, A. K., and L. Tauxe, 1982, Isochronous fluvial systems in


Miocene deposits of northern Pakistan, Sedimentology, v. 29, p.
331-352.

Bender, F., 1975, Geology of the Arabian Peninsula: Jordon, U. S. Geol.


Survey Prof. Paper 560-1, 36 pp.

Boyer, S. E., and D. Elliott, 1982, Thrust systems, Amer. Assoc.


Petrol. Geol. Bull., v. 66, p. 1196-1230.

Burbank, D. W., 1982, The chronologic and stratigraphic evolution of


the Kashmir and Peshawar intermontane basins, northwestern
Himalaya, Ph. D. thesis, Dartmouth College, Hanover, New Hamp-
shire, 291 pp.

Burbank, D. W., 1983, The chronology of intermontane-basin development


in the northwestern Himalaya and the evolution of the Northwest
Syntaxis, Earth Planet. Sci. Lett., v. 64, p. 77-92.

Burbank, D. W., and R. G. Raynolds, in prep., Stratigraphic keys to the


timing of deformation: an example from the northwestern Himalayan
foredeep.

Burbank, D. W., R. G. Raynolds, and G. D. Johnson, 1986, Late Cenozoic


tectonics and sedimentation in the northwestern Himalayan fore-
91

deep: II. Eastern limb of the northwest syntaxis and regional


synthesis, in P. A. Allen, and Homewood, P., eds., Foreland
basins, Spec. Publs. Int. Assoc. Sediment., v. 8, p. 293-306.

Burbank, D. W., R. G. Raynolds, and D. P. Whistler, in press, Detailed


chronologies of tectonic rotations: new insights from magnet-
ostratigraphic studies in Pakistan and along the Garlock fault.

Burbank, D. W., and R. A. K. Tahirkheli, 1985, The magnetostratigraphy,


fission-track dating, and stratigraphic evolution of the Peshawar
intermontane basin, northern Pakistan, Geol. Soc. Amer. Bull.,
v. 96, p. 539-552.

Butler, R. H. W., 1982, The terminology of structures in thrust belts,


Jour. Struc. Geol., v. 4, p. 239-245.

Butler, R. W. H., M. P. Coward, G. M. Harwood, and R. J. Knipe, in


press, Salt, its control on thrust geometry, structural style and
gravitational collapse along the Himalayan mountain front in the
Salt Range of northern Pakistan.

Charlesworth, H. A. K., and W. E. Kilby, 1981, Thrust nappes in the


Rocky Mountain foothills near Mountain Park, Alberta, in K. R.
McClay, and N. J. Price, eds., Thrust and Nappe Tectonics, Geol.
Soc. London Spec. Pub. 9, p. 475-482.

Chevren, V. B., 1986, Tethys-marginal sedimentary basin in western


Iran, Geol. Soc. America Bull., v. 97, p. 516-522.

Cooper, M. A, 1983, The calculation of bulk strain in oblique and


inclined balanced cross-sections, Jour. of Struc. Geol., v. 5,
p. 161-165.

Cooper, M. A., and P. M. Trayner, 1986, Thrust-surface geometry:


implications for thrust belt evolution and section balancing
techniques, Jour. Struc. Geol., v. 8, p. 305-312.

Cotter, G. P., 1933, The geology of the part of the Attock District
west of longitude 72° 45', Geol Soc. India Mem., v. 55, p. 63-156.

Coward, M. P., 1982, Surge zones in the Moine thrust zone of NW


Scotland, Jour. Struc. Geol., v. 4, p. 247-256.

Coward, M. P., and R. W. H. Butler, 1985, Thrust tectonics and the deep
structure of the Pakistan Himalaya, Geology, v. 13, p. 417-420.

Coward, M. P., and J. H. Kim, 1981, Strain within thrust sheets, in


K. R. McClay, and N. J. Price, eds., Thrust and nappe tectonics,
Geol. Soc. London Spec. Pub. 9, p. 275-292.

Cox, A., 1969, Geomagnetic reversals, Science, v. 163, p. 237-244.

Crawford, A. R., 1974, The Salt Range, the Kashmir syntaxis and the
92

Pamir arc, Earth Planet. Sci. Lett., v. 22, p. 371-379.

Dahlstrom, C. D. A., 1969, Balanced cross-sections, Can. Jour. Earth


Sci., v. 6, p. 743-757.

Dahlstrom, C. D. A., 1977, Structural Geology in the eastern margin of


the Canadian Rocky Mountains, in E. L. Heisey, E. R. Norwood, P.
H. Wach, and L. A. Hale, eds., Wyoming Geol. Assoc. 29th Ann.
Field Conf. Guidebook, p. 407-439.

Davis, D. M., and T. Engelder, 1985, The role of salt in fold-and-


thrust belts, Tectonophysics, v. 119, p. 67-88.

Davis, D. M., J. Suppe, and F. A. Dahlen, 1983, Mechanics of fold-and-


thrust belts and accretionary wedges, Jour. Geoph. Res., v. 88,
p. 1153-1172.

Dubey, A. K., 1980, Model experiments showing simultaneous development


of folds and transcurrent faults, Tectonophysics, v. 65, p. 69-84.

Duroy, Y., 1986, Subsurface densities and lithospheric flexure of the


Himalayan foreland in Pakistan from gravity and seismic reflection
constraints, M. S. thesis, Oregon State University, Corvallis,
Oregon, 74 pp.

Duroy, Y., A. Farah, L. L. Malinconico, and R. J. Lillie, 1987,


Subsurface densities and lithopheric flexure of the Himalayan
foreland in Pakistan, in L. L. Malinconico, and R. J. Lillie,
eds., Tectonics and geophysics of the western Himalaya, Geol.
Soc. Amer., Spec. Paper.

Eisbacher, G. H., M. A. Carrigy, and R. B. Campbell, 1974, Paleodrain-


age pattern and late orogenic basins of the Canadian Cordillera,
Soc. Econ. Paleon. and Min., Spec. Pub. v. 22, p. 143-166.

Elliott, D., 1976, The energy balance and deformation mechanisms of


thrust sheets Royal Soc. London Philos. Trans., ser. A, v. 283,
p. 289-312.

Elliott, D., 1983, The construction of balanced cross-sections, Jour.


Struc. Geol., v. 5, p. 101.

Faill, R. T., 1973, Kink band folding, Valley and Ridge Province,
Pennsylvania: Geol. Soc. America Bull., v. 4, p. 1289-1314.

Farah, A., M. A. Mirza, M. A. Ahmad and M. H. Butt, 1977, Gravity field


of the buried shield in the Punjab plain, Pakistan, Geol. Soc.
Amer. Bull., v. 88, p. 1147-1155.

Finetti, I., F. Giorgetti and G. Poretti, 1979, The Pakistani segment


of the DSS profile, Nanga Parbat-Kabul (1974-1975), Bollettino di
Geofisica Teorica ed Applicata, v. 21, p. 159-171.
93

Fischer, M. W., and M. P. Coward, 1982, Strains and folds within thrust
sheets: an analysis of the Heilam Sheet, Tectonophysics, v. 88, p.
291-312.

Fox, F. G., 1983, Structure sections across Parry Islands fold belt,
and Vesey Hamilton Salt Wall, Arctic Archipelago, Canada, in
Seismic Expression of Structural Styles, Amer. Assoc. Petr. Geol.,
Stud. Geol. 15, Vol. 3, A. W. Bally ed., pp. 3.4.1-54 3.4.1-72.

Gardezi, A. H., and M. Asraf, 1974, Gravitative structures of the


Kattha Masral region of the central Salt Range, Pakistan, The
geological bulletin of the Punjab University, n. 11, p. 75-80.

Gee, E. R., 1934, The Saline Series of northwest India, Curr. Sci. 2,
460 p.

Gee, E. R., 1980, Salt Range series geological maps, 1:50,000, 6


sheets, Directorate of Overseas Surveys, United Kingdom, for
Govt. of Pakistan and Pakistan Geol. Surv.

Gee, E. R., 1983, Tectonic problems of the Sub-Himalayan region of


Pakistan, Kashmir Jour. Geol., v. 1, p. 11-18.

Gill, W. D., 1952, The stratigraphy of the Siwalik Series in the


northern Potwar, Punjab, Pakistan, Geol. Soc. London Quart. Jour.,
v. 107, p. 395-421.

Gwinn, V. E., 1964, Thin-skinned tectonics in the Plateau and north-


western Valley and Ridge provinces of the central Appalachians,
Geol. Soc. Amer. Bull., v. 75, p. 863-899.

Hossack, J. R., 1983, A cross-section through the Scandinavian Cale-


donides constructed with the aid of branch line maps, Jour.
Struc. Geol., v. 5, p. 103-111.

Issacs, B., J. Oliver, and L. Sykes, 1968, Seismology and the new
global tectonics, Jour. Geophys. Res., v. 73, p. 5855-5899.

Jaume, S., 1986, Thrust mechanics in the Himalayan foreland of


Pakistan: quantitative and qualitative assessment of an area
underlain by evaporites, M. S. thesis, Oregon State University,
Corvallis, Oregon, 60 pp.

Jaume, S. C., and Lillie, R. J., in press, Active tectonics of the Salt
Range-Potwar Plateau, Pakistan: mechanics of a fold-and-thrust
belt underlain by evaporites, Tectonics.

Jenyon, M. K., 1985, Fault-associated salt flow and mass movements,


Geol. Soc. London, v. 142, p. 547-553.

Johnson, G. D., N. M. Johnson, N. D. Opdyke, and R. A. K. Tahirkheli,


1979, Magnetic reversal stratigraphy and sedimentary tectonic
history of the upper Siwalik Group, eastern Salt Range and
94

southwestern Kashmir, in Geodynamics of Pakistan, A. Farah and


K. A. DeJong, eds., Geol. Surv. Pakistan, p. 149-165.

Johnson, G. D., R. G. H. Raynolds and D. W. Burbank, 1986, Late Ceno-


zoic tectonics and sedimentation in the northwestern Himalayan
foredeep: I. Thrust ramping and associated deformation in the
Potwar region, in P. A. Allen, and P. Homewood, eds., Foreland
basins, Spec. Pubis. Int. Assoc. Sediment., v. 8, p. 273-291.

Johnson, N. M., Opdyke, N. D., Johnson, G. D., Lindsay, E. H., and


Tahirkeli, R. A. K., 1982, Magnetic polarity stratigraphy and ages
of Siwalik Group rocks of the Potwar Plateau, Pakistan, Palaeo-
geog. Plaeoclim. Palaeoecol., v. 37, p. 17-42.

Jones, P. B., 1982, Oil and gas beneath east-dipping underthrust faults
in the Alberta foothills, Rocky mountain Assoc. Geol. Guidebook,
Denver, Colorado, p. 61-74.

Kazmi, A. H., and R. A. Rana, 1982, Tectonic map of Pakistan, Geol.


Surv. of Pakistan, scale 1:2,000,000, 1st. edition.

Keller, H. M., R. A. K. Tahirkheli, M. A. Mirza, G. D. Johnson, N. M.


Johnson, and N. M. Opdyke, 1977, Magnetic polarity stratigraphy of
the Upper Siwalik deposits, Pabbi Hills, Pakistan, Earth Planet.
Sci. Lett., v. 36, p. 187-201.

Khan, M. J., 1984, Brief results of the paleomagnetic studies of the


Siwalik Group of the Trans-Indus Salt Range, Pakistan, Geol. Bull.
Univ. Peshawar, v. 17, p. 176-178.

Khan, M. A., R. Ahmed, H. A. Raza, and A. Kemal, 1986, Geology of


Petroleum in Kohat-Potwar Depression, Pakistan, Amer. Assoc. Pet.
Geol. Bull., v. 70, p. 396-414.

Klootwijk, C. T., 1979, A summary of palaeomagnetic data from extra-


peninsular IndoPakistan and south central Asia: implications for
collision tectonics, in P. S. Saklani, ed., Structural geology of
the Himalaya: Today and Tomorrow Printers and Pub., New Delhi, p.
307-360.

Knipe, R. J., 1985, Footwall geometry and the rheology of thrust


sheets: Jour. Struc. Geol., v. 7, p. 1-10.

Latif, M. A., 1973, Partial extension of the evaporite facies of the


Salt Range to Hazara, Pakistan: Nature, v. 244, p. 124-125.

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, M. S. thesis, Oregon State Univ.,
Corvallis, Oregon.

Lillie, R. J. and M. Yousuf, 1986, Modern analogs for some midcrustal


reflections observed beneath collisional mountain belts, in
95

Reflection Seismology: The Continental Crust, Amer. Geoph. Un.,


Geodynamics Series Vol. 14, M. Barazangi and L. Brown, eds.,
p. 55-65.

Lillie, R. J., G. D. Johnson, M. Yousuf, and R. S. Yeats, 1987,


Structural development within the Himalayan foreland fold-and-
thrust belt of Pakistan, in Basins of Eastern Canada and Worldwide
Analogs, Canada. Assoc. Petrol. Geol., Memoir 12, C. Beaumont and
A. J. Tankard, eds., in press.

Lohmann, H. H., 1972, Salt dissolution in subsurface of British North


Sea as interpreted from seismograms: Am. Assoc. Petrol. Geol.
Bull., v. 56, p. 472-479.

Lyon-Caen, H., and P. Molnar, 1985, Gravity anomalies, flexure of the


Indian plate, and the structure, support and evolution of the
Himalaya and Ganga basin, Tectonics, v. 4, p. 513-538.

Mandl, G., and G. K. Shippan, 1981, Mechanical model of thrust sheet


gliding and imbrication, in K. R. McClay, and N. J. Price,
eds., Thrust and nappe tectonics: Geol. Soc. London Spec. Pub. 9,
p. 79-98.

Martin, N. R., 1962, Tectonic style in the Potwar, West Pakistan:


Punjab Univ. Geol. Bull., v. 2, p. 39-50.

Martinez, J. D., 1974, Tectonic behavior of Evaporites, in Fourth


International Symposium on salt: Northern Ohio Geol. Soc., p.
155-168.

McDougall, J., in prep., Kalabagh strike-slip fault zone, western Salt


Range, Pakistan, PhD thesis, Oregon State University, Corvallis,
in prep.

McDougall, J. W., 1985, Strike-slip faulting in a foreland fold-thrust


belt, western Salt Range, central Pakistan, Geol. Soc. Amer.,
Abs. with Prog., p. 658.

Minster, J. B., and T. H. Jordan, 1978, Present day plate motions,


Jour. Geoph. Res., v. 83, p. 5331-5354.

Mitra, S., 1986, Duplex structures and imbricate thrust systems: geom-
etry, structural position, and hydrocarbon potential, Am. Assoc.
Petrol. Geol. Bull., v. 70, p. 1087-1112.

Mitra, S., and S. Boyer, 1986, Energy balance and deformation mechan-
isms of duplexes, Jour. Struc. Geol., v. 8, p. 291-304.

Molnar, P., B. C. Burchfiel, L. K'uangyi, and Z. Ziyan, 1987, Geomor-


phic evidence for active faulting in the Altyn Tagh and northern
Tibet and qualitative estimates of its contribution to the
convergence of India and Eurasia, Geology, v. 15, p. 249-253.
96

Molnar, P., W. P. Chen, T. J. Fitch, P. Tapponier, W. E. K. Warsi and


F. T. Wu, 1976, Structure and tectonics of the Himalaya: A brief
summary of relevant geophysical observations, Colloques Inter-
national du Centre National de la Recherch Scientific, no. 268,
p. 269-294.

Molnar, P., and Deng Qidong, 1984, Large earthquakes and the average
rate of deformation in Asia, Jour. Geoph. Res., v. 89, p. 6203-
6228.

Molnar, P., and P. Tapponnier, 1975, Late Cenozoic tectonics of Asia:


effects of a continental collision, Science, v. 189, p. 419-426.

Morley, C. K., 1986, Classification of thrust fronts, Am. Assoc.


Petrol. Geol. Bull., v. 70, p. 12-25.

Morse, J. 1977, Deformation in ramp regions of overthrust faults in


E. L. Heisey, D. E. Lawson, E. R. Norwood, P. H. Wach, and L. A.
Hale, eds., Rocky Mountain thrust belt, geology and resources:
Wyoming Geol. Assoc. Guidebook 29, p. 457-470.

Mugnier, J. L., and P. Vialon, 1986, Deformation and displacement of


the Jura cover on its basement: Jour. Struc. Geol., v. 8, p.
373-388.

Murrell, S. A. F., 1981, The rock mechanics of thrust and nappe


formation, in K. R. McClay, and N. J. Price, eds., Thrust and
nappe tectonics, Geol. Soc. London Spec. Pub. 9, p. 99-109.

Ni, J., and M. Barazangi, 1984, Seismotectonics of the Himalayan


collision zone: Geometry of the underthrusting Indian plate
beneath the Himalaya, Jour. Geoph. Res., v. 89, p. 1147-1163.

Opdyke, N. D., N. M. Johnson, G. D. Johnson, E. H. Lindsay, and R. A.


K. Tahirkheli, 1982, The paleomagnetism of the middle Siwalik
formations of northern Pakistan, Paleogeo., Paleoclim., and
Paleoecol., v. 37, p. 43-62.

Opdyke, N. D., E. Lindsay, G. D. Johnson, N. Johnson, R. A. K. Tahirk-


heli, and M. A. Mirza, 1979, Magnetic polarity stratigraphy and
vertebrate paleontology of the Upper Siwalik subgroup of northern
Pakistan: Palaeogeog. Palaeoclim. Palaeoecol., v. 27, p. 1-34.

Pennock, N., in prep., Balanced structural cross-section of the eastern


Potwar Plateau, Pakistan: Shortening and deformation at the edge
of a salt basin, M. S. thesis, Oregon State University, Corvallis.

Pilgrim, G. E., 1910, Preliminary notes on a revised classification of


the Tertiary freshwater deposits of India, Rec. Geol. Surv. India,
v. 40, p. 185-205.

Pilgrim, G. E., 1913, The correlation of the Siwaliks with mammal


horizons of Europe, Rec. Geol. Soc. India, v. 43, p. 264-326.
97

Pinfold, E. S., 1918, Notes on structure and stratigraphy in the


North-west Punjab: Rec. Geol. Survey India, v. 49, p. 137-160.

Powell, C. M., 1979, A speculative tectonic history of Pakistan and


surroundings: Some constraints from the Indian Ocean, in A. Farah
and K. A. DeJong, eds., Geodynamics of Pakistan, Geological Survey
of Pakistan, p. 5-24.

Price, R.A., 1981, The Cordilleran foreland thrust and fold belt in the
southern Canadian Rocky Mountains, in K. R. McClay and N. J.
Price, eds., Thrust and Nappe Tectonics, Geol. Soc. London, Black-
well Scientific Pubs., p. 427-448.

Quittmeyer, R. C., and K. H. Jacob, 1979, Historical and modern seis-


micity of Pakistan, Afghanistan, northwestern India and southeast-
ern Iran, Bull. Seismol. Soc. Amer., v. 69, p. 773-823.

Ramberg, H., 1974, Experimental and theoretical study of salt dome


evolution, in A.H. Coogan, ed., Fourth Symposium on salt, Northern
Ohio Geol. Soc., Cleveland, OH, v. 1, p. 261-270.

Ramsay, J. G., 1969, The measurement of strain displacement in orogenic


belts, in P. E. Kent, G. E. Satterthswaite, and A. M. Spencer,
eds., Time and Place in Orogeny, Geol. Soc. London Spec. Pub. 3,
p. 43-79.

Ramsay, J. G., M. Casey, and R. Kligfield, 1983, Role of shear in


development of the Helvetic fold-thrust belt of Switzerland,
Geology, v. 11, p. 439-442.

Rayleigh, C. B., and D. T. Griggs, 1963, Effect of the toe in the


mechanics of overthrust faulting, Geol. Soc. Amer. Bull., v. 74,
p. 819-830.

Raynolds, R. G., 1980, The Plio-Pleistocene structural and strati-


graphic evolution of the eastern Potwar Plateau, Pakistan,
Ph. D. thesis, Dartmouth College, Hanover, New Hampshire, 265 pp.

Raynolds, R. G., 1981, Did the ancestral Indus flow into the Ganges
drainage?, Geol. Bull. Univ. Peshawar, v. 14, p. 141-150.

Raynolds, R. G., and G. D. Johnson, 1985, Rates of Neogene depositional


and deformational processes, north-west Himalayan foredeep margin,
Pakistan, in N. J. Snelling, ed., The Chronology of the Geological
Record, Geol. Soc. London, Memoir No. 10, p. 297-311.

Rich, J. L., 1934, Mechanics of low-angle overthrust faulting as


illustrated by the Cumberland thrust block, Virginia, Kentucky,
Tennessee, Amer. Assoc. Petrol. Bull., v. 18, p. 1584-1596.

Rubey, W. W., and M. K. Hubbert, 1959, Role of fluid overpressure in


mechanics of overthrust faulting, Geol. Soc. Amer. Bull., v. 70,
98

p. 167-206.

Sarwar, G., and K. A. DeJong, 1979, Arcs, oroclines, syntaxes: the


curvatures of mountain belts in Pakistan, in Geodynamics of
Pakistan, A. Farah and K. A. DeJong, eds., Geol. Surv. of
Pakistan, Quetta, p. 341-349.

Shroder, J. F., in prep., Quaternary history of middle and upper Indus


and surrounding valleys, northern Pakistan.

Seeber, L., and J. G. Armbruster, 1979, Seismicity of the Hazara Arc in


northern Pakistan: decollement vs. basement faulting, in A. Farah
and K. A. DeJong, eds., Geodynamics of Pakistan, Geological Survey
of Pakistan, p. 131-142.

Seeber, L., J. G. Armbruster and R. C. Quitmeyer, 1981, Seismicity and


continental subduction in the Himalayan Arc, in H. K. Gupta and
F. M. Delany, eds., Zagros, Hindu Kush, Himalaya Geodynamic Evolu-
tion, Amer. Geoph. Union, Geodynamics Series 3, p. 215-242.

Seeber, L., and K. H. Jacob, 1977, Microearthquake survey of northern


Pakistan, preliminary results and tectonic implications, Proc.
CNRS Symp. Geol. Ecology Himalayas, Paris, p. 347-360.

Serra, S., 1977, Styles of deformation in the ramp regions of over-


thrust faults, in E. L. Heisey, E. R. Norwood, P. H. Wach, and
L. A. Hale, eds., Wyoming Geol. Assoc. 29th Ann. Field Conf.
Guidebook, p. 487-498.

StOcklin, J., 1974, Possible ancient continental margin in Iran: in


C. L. Drake, and C. A. Burk, eds., The Geology of Continental
Margins, Springer-Velag, New York, p. 889-903.

Suppe, J., 1983, Geometry and kinematics of fault-bend folding, Amer.


Jour. Sci., v. 283, p. 684-721.

Suppe, J., 1985, Principles of structural geology: Englewood Cliffs,


N. J., Prentice-Hall, 537 p.

Swaminath, J., V. Venkatesh, and R. K. Sundaram, 1964, Role of


Precambrian lineaments in the evolution of Cenozoic festoons of
the Indian subcontinent, Internat. Geol. Cong. 22nd, New Delhi,
1964, sec. 11, p. 316-333.

Tapponnier, P., and P. Molnar, 1977, Active faulting and tectonics in


China, Jour. Geophys. Res., v. 82, p. 2905-2943.

Tapponnier, P., G. Peltzer, A. Y. Le Dain, R. Armijo, and P. Cobbold,


1982, Propagating extrusion tectonics in Asia: new insights from
simple experiments with plasticine, Geology, v. 10, p. 611-616.

Tauxe, L., and N. D. Opdyke, 1982, A time framework based on magnet-


stratigraphy for Siwalik sediments of the Khaur area, northern
99

Pakistan, Palaeogeog., Palaeoclim., Palaeoecol., v. 37, p. 43-61.

Telford, W. M., L. P. Geldart, R. E. Sheriff, and D. A. Keys, 1976,


Applied geophysics, Cambridge University Press, 860 p.

Thompson, R. I., 1981, The mature and significance of large "blind"


thrusts within the northern Rocky mountains of Canada, in K. R.
McClay, and N. J. Price, eds., Thrust and Nappe Tectonics, Geol.
Soc. London Spec. Pub. 9, p. 449-462.

Vann, I. R., R. H. Graham, and A. B. Hayward, 1986, The structure of


mountain fronts, Jour. Struc. Geol., v. 8, p. 215-228.

Voskresenskiy, I. A., 1978, Structure of the Potwar highlands and Salt


Range as a indicator of horizontal movement of the Hindustan
Platform, Geotectonics, v. 12, p. 70-75.

Wadia, D. N., 1928, The geology of the Poonch State (Kashmir) and
adjacent portions of the Punjab, Mem. Geol. Survey of India, v.
51, p. 185-370.

Wadia, D. N., 1931, The syntaxis of the north-west Himalaya: its rocks,
tectonics and orogeny, Rec. Geol. Surv. India, v. 65, p. 189-220.

Wells, N. A., 1984, Marine and continental sedimentation in the early


Cenozoic Kohat Basin and adjacent northwestern Pakistan, Ph.D.
dissertation, University of Michigan, Ann Arbor, 465 pp.

Williams, G., and Chapman, T., 1983, Strains developed in the hanging-
walls of thrusts due to their slip/propagation rate: a dislocation
model, Jour. Struc. Geol., v. 5, p. 563-571.

Wiltschko, D. V., 1979, A mechanical model for thrust sheet deformation


at a ramp, Jour. Geophys. Res., v. 84, p. 1091-1104.

Wiltschko, D. V., and W. M. Chapple, 1977, Flow of weak rocks in Appa-


lachain Plateau folds, Am. Assoc. Petrol. Geol. Bull., v. 61,
p. 653-670.

Wiltschko, D., and D. Eastman, 1983, Role of basement warps and faults
in localizing thrust fault ramps, in R. D. Hatcher, H. Williams,
and I. Zietz, eds., Contributions to the Tectonics and Geophysics
of Mountain Chains, Mem. 158, p. 177-190.

Windley, B. F., 1983, Metamorphism and tectonics of the Himalaya: Jour.


Geol. Soc. Lon., v. 140, p. 849-865.

Woodward, N. B., S. E. Boyer, and J. Suppe, 1985, An outline of


balanced cross-sections, notes for a short course on balanced
sections sponsored by the Structural Geology and Tectonics Divi-
sion of the Geol. Soc. America at annual meeting, Orlando, Fl.,
Univ. of Tenn. Dept. of Geol. Sci. Studies in Geology 11, 170 pp.
100

Yeats, R. S., in prep., Active faults related to folding.

Yeats, R. S., and A. Hussain, in press, Timing of structural events in


the Himalayan foothills of northwestern Pakistan, in prep.

Yeats, R. S., S. H. Khan and A. Akhtar, 1984, Late Quaternary deforma-


tion of the Salt Range of Pakistan, Geol. Soc. Amer. Bull., v. 95,
p. 958-966.

Yeats, R. S., and R. D. Lawrence, 1984, Tectonics of the Himalayan


thrust belt in northern Pakistan, in B U. Haq, and J. D. Milliman,
eds., Marine Geology and Oceanography of Arabian Sea and Coastal
Pakistan, p. 177-198.

Yeats, R. S., T. Nakata, K. S. Valdiya, M. Pandey, A. Farah, M. A.


Mirza, M. Fort, and R. Armijo, in prep., Active fault map of the
Himalaya, 1:2,000,000, IGCP 206 Atlas of Worldwide Active Faults.

Zharkov, M. A, 1981, History of Paleozoic salt accumulation: Berlin,


Springer-Velag, 308 p.
APPENDICES
101

APPENDIX A

DEVELOPMENT OF THE HIMALAYAN FRONTAL THRUST ZONE:

SALT RANGE, PAKISTAN

Dan M. Baker, Robert J. Lillie, and Robert S. Yeats


Department of Geology
Oregon State University
Corvallis, Oregon, 97331

Gary D. Johnson
Department of Earth Sciences
Dartmouth College
Hanover, New Hampshire, 03755

Mohammad Yousuf and Agha Sher Hamid Zamin


Oil and Gas Development Corporation
13-N, Markaz F/7
Islamabad, Pakistan

CONTENTS
Abstract 102
Introduction 102
Regional Setting 102
Structure of the Salt Range/Potwar Plateau 104
Timing of Ramping 108
Discussion 111
Ramping 111
Calculation of Shortening Rate 111
Conclusions 114
Acknowledgements 115

Submitted to Geology
April 10, 1987
102

ABSTRACT

The Salt Range is the active frontal thrust zone of the Himalaya in
Pakistan. Seismic reflection data show that a 1 km offset of the
basement acted as a buttress that caused the central Salt Range/Potwar
Plateau thrust sheet to ramp to the surface, exposing Mesozoic and
Paleozoic strata at the range front. The frontal portion of the thrust
sheet was folded passively as it overrode the subthrust surface on a
ductile layer of salt. The hindward portion of the thrust sheet is
undeformed, except for minor folding. Lack of internal deformation is
due to decoupling of sediments from the basement along this Eocambrian
salt layer.
The timing of ramping is constrained by stratigraphic evidence,
paleomagnetic dating of unconformities and sediment accumulation
rates. It is estimated that the thrust sheet overrode the basement
offset from 2.1 to 1.6 mya. Cross-section balancing gives a minimum of
19 to 23 km of shortening across the ramp. The rate of Himalayan
convergence which can be attributed to underthrusting of Indian
basement beneath sediments in the Pakistan foreland is therefore 9-14
mm/yr.

INTRODUCTION

Approximately 3000 km of seismic reflection profiles (supplied by


the Pakistan Oil and Gas Development Corporation and Ministry of
Petroleum), along with surface maps and well data, have been used to
construct balanced cross-sections of the foreland fold-thrust belt of
the Himalaya in Pakistan. The Salt Range, the active thrust front,
provides a modern analogue to extinct thrust fronts and, to a lesser
degree, internal regions of fold-thrust belts.
Recent sedimentological and stratigraphic studies in the Potwar
Plateau/Salt Range have used magnetostratigraphy and tephrochronology
to temporally constrain horizons in the Siwalik molasse (e.g. Johnson
et al., 1986). These data constrain the timing of deformation and the
progression of the deformation front across the foreland basin
(Raynolds and Johnson, 1985). The rate of shortening at the front of
the fold-thrust belt can be determined by using this timing information
along with estimates of contraction of the sedimentary cover based on
the balanced cross-sections. Ramping of the Salt Range thrust sheet
represents a discrete event from which timing and contraction can be
determined. The resulting shortening rate can then be used to infer
the portion of Indian-Asian plate convergence attributable to
underthrusting of the Indian basement beneath sediments at the Salt
Range front.

REGIONAL SETTING

The southern margin of the Himalayan collision zone in Pakistan is


the Salt Range and Kohat/Potwar Plateau, an active foreland fold-thrust
belt formed in response to the underthrusting of cratonic India beneath
its own Phanerozoic sedimentary cover (fig. Al). At the Hazara-Kashmir
Syntaxis, the fold-thrust belt changes in trend and broadens in
103

Fig. Al: Position of the Salt Range Thrust front and Potwar Plateau
(enclosed in box) within the overall Himalayan trend. HKS=Hazara-
Kashmir Syntaxis; ISZ=Indus Suture Zone; MBT=Main Boundary Thrust;
MFT=Main Frontal Thrust (active); MKT=Main Karakorum Thrust; SH=
Sargodha High. Rectangular box denotes area of figure 2.
104

Pakistan. In India west of the Punjab reentrant, there is strong


coupling between sediments and basement; consequently the zone of
underthrusting is narrow and has a high angle of cross-sectional taper
(Davis et al., 1983; Davis and Engelder, 1985). In Pakistan, on the
west flank of the Hazara-Kashmir Syntaxis, low-strength evaporites of
late Precambrian to Early Cambrian age constitute the zone of
decollement (Crawford, 1974). As a result and in contrast to India,
the zone of underthrusting extends far out over the foreland (Seeber
and Armbruster, 1979), more than 100 km south of the Main Boundary
Thrust, and it has a narrow cross-sectional taper. Further evidence
for the decoupling properties of the evaporites lies in the large
detachment earthquakes which occur beneath the foreland in India, but
not in Pakistan (Quittmeyer and Jacob, 1979; Seeber et al., 1981).
In northern Pakistan, a southward progression of major deformation
has been documented (Raynolds and Johnson, 1985; Johnson et al., 1986;
Burbank et al., 1986). At the southern margin of the Peshawar basin,
major thrusting occurred prior to 2.8 mya (Burbank, 1982). Farther
south, in the northern Potwar Plateau, major deformation ended by 1.9
mya (Raynolds and Johnson, 1985). In the eastern Salt Range, major
deformation may be younger than 0.4 mya (Johnson et al., 1979). The
deformation front in the central Salt Range may now be shifting 15 km
south to the Lilla structure, a thrust-cored anticline in the Jhelum
River plain (fig. A2). In India, Acharyya and Ray (1982) showed that
progression of the thrust front is accompanied by southward migration
of the Siwalik molasse depositional axis. Likewise, Raynolds and
Johnson (1985) estimate that the depocenter in Pakistan is migrating
southward at 30 mm/yr, and is currently located just south of the Salt
Range, along the Jhelum River.
South of the fold-thrust belt, sediments slope gently northward from
the Sargodha High, an area of moderate seismicity including exposures
of Precambrian rocks (Farah et al., 1977). This ridge is parallel to
the overall northwest-southeast Himalayan trend and may be related to
downflexing of the Indian plate (Seeber et al., 1981; Yeats and
Lawrence, 1984; Duroy et al., 1987).

STRUCTURE OF THE SALT RANGE/POTWAR PLATEAU

At the Salt Range front, Eocambrian evaporites and overlying strata


override synorogenic fan material and alluvium (Yeats et al., 1984).
The central Salt Range is strongly emergent, located between a weakly
emergent thrust front at the Surghar Range (J. McDougall, in prep) and
a buried thrust front in the easternmost Salt Range (terminology of
Morley, 1986). The Salt Range lies approximately 80 km outboard of
thrusting in the northern Potwar Plateau; the intervening Soan trough
is relatively undeformed (fig. A3). Seismic-reflection profiles in the
region (Khan et al., 1986; Lillie et al., in press) show that the Salt
Range and southern Potwar Plateau constitute a large slab that is being
thrust over the foreland with very little internal deformation.
The northern flank of the Salt Range is an eroded monocline that
flattens northward into the southern limb of the Soan trough; the
monocline is the surface expression of the footwall ramp. Reflection
profiles show that lower Siwalik sediments exposed at the surface
overlie a fault with a basement offset of one kilometer, down-to-the
105

Fig. A2: Geologic map (after Gee, 1980) showing locations of structur-
al features mentioned in text. HR=Hill Ranges; KF=Kalabagh Fault;
L=Lilla Anticline; NPDZ=Northern Potwar Deformed Zone; PH=Pabbi
Hills Anticline; RF=Riwat Fault; SRFT=Salt Range Frontal Thrust.
Cities: I=Islamabad; R=Rawalpindi. Exploration wells: D=Dhariala;
L=Lilla. Cross-section traces: A-A' (fig. A3); C-C' (fig. A5),
seismic line: B-B' (fig. A4). Magnetostratigraphic sample loca-
tions: BN=Baun; KK=Kotal Kund. Solid lines represent structures
mapped from Landsat photographs, dashed lines are less prominent
features, and dotted lines denote buried structures.
72° 73°

t
ALLUVIUM, SIWALIK,and CAMBRIAN to CRETACEOUS
RAWALPINDI MOLASSE 0 50km
* MAGNETOSTRATIGRAPHIC SITES
111111111
PALEOCENE and EOCENE EOCAMBRIAN SALT RANGE FM. -EXPLORATION WELLS
Figure A2
N. SOAN S. SALT
POTWAR SYN. POTWAR RANGE C

/1 //\ 11-/ H--


10 I 10

20
H TERTIARY MOLASSE
2
FrA CAMBRIAN TO' EOCENE
EOCAMBRIAN 30

PRECAMBRIAN BASEMENT
,10 KM V.E. 1.1
40
MOHO

50
100 80 60 40 20 0 -20
DISTANCE FROM CENTER OF SALT RANGE (KM)

Fig. A3: Schematic cross-section of the Himalayan foreland fold-thrust


belt in Pakistan (see Lillie et al., in press, for seismic
profiles used to construct section). See figure 1 for location.
108

north (Lillie and Yousuf, 1986; fig. A4). This ramp and the underlying
basement offset correlate with a local steepening of the Bouguer
gravity gradient (Farah et al., 1977). The basement offset acted as a
buttress which controlled ramping and emplacement of the Salt Range
thrust sheet. Passive folding, as the thrust sheet cut upsection, then
flattened, is primarily responsible for the broad anticlinal structure
of the central Salt Range. The Salt Range, like most anticlines of the
Potwar Plateau, is cored by salt which appears to have flowed from
beneath synclines (cf. Davis and Engelder, 1985; Fox, 1983), as
suggested by the presence of over 2 km of salt at the bottom of the
Dhariala well (Gee, 1983; fig. A2).
In order to control ramping, the basement offset must have been
present prior to the encroachment of the thrust front. Hypotheses for
the origin of the down-to-the north basement fault include: (a)
extension related to Eocambrian rifting that may have accompanied the
formation of evaporite basins on the northwestern margin of the Indian
subcontinent, or Mesozoic rifting apart of Gondwana; (b) reverse
faulting related to either the latest Cretaceous through Paleocene
convergence in the Attock-Cherat Range or the present compressional
regime (Yeats and Hussain, in press); or (c) Neogene normal faulting
related to flexure as the crust is loaded by thrust sheets from the
north (Lillie and Yousuf, 1986; Duroy et al., 1987). Stratigraphic
relations suggest that hypothesis (c) is the most likely.
The right-lateral Kalabagh tear fault, terminates the Salt Range
to the west (Yeats and Lawrence, 1984). In contrast, the eastern
termination of the Salt Range is divided into several fault blocks
bounded by forward- and hindward-verging thrusts (Johnson et al.,
1986). Farther to the north, in the eastern Potwar Plateau, folds and
thrusts are aligned en echelon in a SSW-NNE direction. The increase in
deformation from the central to the eastern Potwar Plateau may be the
result of differential drag at the edge of the salt basin caused by the
contrast in basal friction as the detachmentzone salt facies thins
eastward (Davis and Engelder, 1985) and as the dip on the basement
decreases (Jaume and Lillie, in press).

TIMING OF RAMPING

The timing of thrusting is often bracketed by the youngest rocks


involved in deformation and the oldest sediments overlapping these
deformed strata. Yeats and others (1984) documented that the Salt
Range thrust is young and possibly still active; the initiation of
ramping, however, is more difficult to constrain. In the central Salt
Range at the Baun locality (Fig. A2), the top of the Dhok Pathan
formation, dated as 5.1 mya or slightly younger (Johnson et al., 1982),
is overlain disconformably by Upper Siwalik sediments with a
conglomerate layer at its base. Dhok Pathan beds are tilted northward
and truncated above the footwall ramp (fig. A4), indicating ramping
began later than 5.1 Ma. In addition, the Upper Siwaliks are
essentially unrotated, while Dhok Pathan beds are rotated 35°
counterclockwize (Opdyke et al., 1982).
An Upper Siwalik conglomerate bed, deposited about 4.5 mya at the
Baun locality, contains clasts of Eocene carbonates and recycled
Precambrian clasts from the Permian Tobra Formation (Johnson et al.,
11111111i I I I II Ill II 11111 1111111111111111111111111111111111111
110

1986; Burbank and Raynolds, in press). These clasts are important


stratigraphic evidence constraining the timing of either ramping or
basement faulting. If the source area of Upper Siwalik clasts is the
Salt Range, then conglomerate deposition in combination with diminished
sedimentation rates and differential rotation may indicate incipient
Salt Range thrusting by 4.5 mya (Johnson et al., 1986; Burbank and
Raynolds, in press). Alternatively, Eocene and Permian clasts are
interpreted as derived from the south, from the up-thrown basement
fault block. Diminished sedimentation rates, erosion and basement
faulting may have occurred when this portion of the crust was at the
crest or along the northern flank of the peripheral bulge (now
underneath the Sargodha High; fig. Al). This interpretation is
supported by the Shell #1 Lilla well (fig. A2), which penetrated
Cambrian rocks directly beneath Siwaliks. Erosion of intervening
strata, including Eocene carbonates and the Lower Permian Tobra
Formation, is therefore demonstrated south of the basement offset. If
Upper Siwalik conglomerates preserved at Baun were derived from the
up-thrown basement fault block, then the down-to-the north, basement
offset (figs. A3, A4) was formed or enhanced by Neogene faulting
related to flexure, as suggested by Lillie and Yousuf (1986; see also
Duroy et al., 1987).
The age of thrust ramp initiation in the central Salt Range
appears poorly constrained by stratigraphic evidence alone. Initiation
of ramping predates deposition of the Potwar Silt, which is believed to
have been ponded behind the rising Salt Range. Raynolds (1980) found
only normal magnetization in the Potwar Silt, indicating a Brunhes age
of less than 0.7 Ma, while Shroder (in press) cites evidence suggesting
that the silt is only as old as 170,000 yr.
To better resolve the onset of ramping, information on the timing of
specific events has been compiled from the eastern margin of the Potwar
Plateau/Salt Range, where deformation is more evenly distributed,
rather than concentrated at the frontal thrust (Johnson et al., 1986;
Burbank et al., 1986). As a result, a more complete record of
sedimentation is exposed there, where the progressive
south-southeastward onset of deformation can be traced from the Riwat
fault to the Pabbi Hills anticline (Johnson et al., 1986; fig. A2).
The age of initiation of folding within the central portion of the Soan
trough is poorly constrained. In the ramp region to the south, initial
deformation may have been expressed between 3.0 and 2.2 mya, by a
change in sediment accumulation throughout the eastern Potwar
Plateau. A major structural event then occurred as the Salt Range
thrust sheet overrode the basement offset between 2.1-1.6 mya. The
timing of this event is suggested by the abrupt termination of folding
in the northern Potwar Plataeu at about 1.9 mya, termination of sedi-
mentation over folds in the eastern Potwar, cessation of sedimentation
in the Kotal Kund region at 1.6 mya (Johnson et al., in press) and
accelerated rotation of the Soan syncline after 2 Ma (Burbank et al.,
in press). Initiation of ramping between 2.1 and 1.6 mya also fits the
chronology of the southeastward-propagating deformation front when the
position of the ramp is projected along strike to the east. At this
time, there was a transfer of strain release from the northern Potwar
Plataeu to the Salt Range ramp, implying the Soan trough/Salt Range
region began to move as one coherent thrust sheet.
111

DISCUSSION

RAMPING: In Fig. A5, the Salt Range is interpreted as a hanging-wall


ramp anticline. The shape of the central Salt Range differs from a
typical fault-bend fold due to its lack of a frontal, forward dipping
limb and the presence of thickened salt at the base of the thrust
sheet. A frontal limb forms as a thrust flattens after cutting up
section (Suppe, 1983). Therefore, a plausible explanation is that the
Salt Range thrust does not flatten substantially, and has a relatively
smooth trajectory. The gentle slope to the surface from the basement
offset is consistent with the thrust cutting upsection through
relatively incompetent molasse, and the thrust sheet climbing over its
own debris.
Loading by the thick molasse sequence within the Soan trough is
believed to cause ductile flow of salt toward the south (Gee, 1983).
The basement fault acting as a buttress may have localized the growth
of an elongate salt pillow or dome as the overlying competent strata
buckled against it. Noting Wiltschko and Eastman's (1983) work on
potential thrust fault orientations associated with basement faults, it
is likely that the normal fault plane would be reactivated during
compression in strata overlying the basement (fig. A5). As salt growth
continued, Cambrian and younger strata on the north side of the fault
would be lifted until they could be pushed across the basement offset.
During this later process, the steep initial ramp angle would be
abandoned in favor of a shallower trajectory in conjunction with salt
thickening in the lower ramp region of the upper plate (eg. mode 2 of
Serra, 1977). This mechanism explains the lack of footwall deformation
normally associated with smooth thrust trajectories (see Cooper and
Trayner, 1986) and thickened salt in the allochthon (e.g. Dhariala
well; Gee, 1983), while avoiding an unreasonably sharp initial
interlimb ramp angle. Therefore, the southern flank of the central
Salt Range is not only an eroded thrust fault scarp (Yeats et al.,
1984), but may also be close to the preserved portion of the downthrown
normal fault block (fig. A5).
Due to its limited lateral extent (fig. A5), the basement offset
does not satisfactorily explain the emplacement of the entire Salt
Range. A thrust fault surface will propagate along strike into areas
where there is no other mechanical cause for failure (Wiltschko and
Eastman, 1983). In the western Salt Range, ramping was probably
facilitated by a basement warp interpreted by Leathers (1987) as the
northeast flank of the Sargodha basement ridge converging with the
overthrust belt. In the eastern Salt Range, a southward thinning of
the detachment zone salt facies may have aided thrust ramping.
Therefore, it appears that a combination of ramping mechanisms was
responsible for the emplacement of the entire Salt Range.

CALCULATION OF SHORTENING RATE: In the western Himalaya at longitude


73° east, the northward motion of the Indian plate relative to the
Asian plate is calculated as 42 mm/yr (Minster and Jordan, 1978).
Convergence is taken up by: (1) shortening of sedimentary strata in the
foreland as India underthrusts its cover rocks; (2) contraction within
the collision zone via uplift in the high Himalaya and reactivation of
112

Fig. A5: Balanced cross-section and restorations demonstrating a


possible mode of origin of the Salt Range. Formation contacts are
dashed above upward limit of preserved strata. Erosion occurred
during Salt Range thrusting. (a) Stratigraphic sequence, at the
end of platform deposition and prior to molasse deposition
(approx. 35 mya). The thinner Cambrian to Eocene sequence south
of future basement offset is interpreted to result from erosion
during subsequent basement faulting. (b) restoration after bulk
of molasse deposition, prior to compressional deformation (approx.
2 to 5 mya). Normal fault formed as Indian lithosphere flexed
downward due to thrust emplacement to the north. (c) Present
configuration. The thrust sheet encountered the basement offset
and ramped above a thickening layer of salt. Salt has flowed into
this section from the north. Molasse removed from the autochthon
during thrusting is not shown. Dashed detachment surface is
schematic. Vertical faults were mapped as "normal or probably
normal" by Gee (1980).
/
I

I
/ / /
k

\
/
I-
/11 /
I 7.

/,
\

\ I/
\ ./ I

, /
I//11
".
c' \/.\ /

\
I

/,
' -1- \11/1 1 -;
\ I I \I \

/-
/I 'I

/\,\A1/1\/\ "\'\-/\JI;7\2/\'\,\\,/,)I /\\1\\__A

/\ /, \,i\/\/\ 7,\ i_\ ,i\i \\/ '


I\-
\ /\ \ I-\/ 1

s,
1\1,/\)\;1,\--71\ /,\
I
\ / \I \/\--i /.'/---
\ \/- I / /-\/ /1 \ _\ I
\/0 /(\/
/ / I \I; \I /
I ,
\ I \
./1\ 1. _\1 1-- BI
0 00 1
o 0 01 POST SIWALIK
SIWALIK TERTIARY
MOLASSE
RAWALPINDI
..... I 1i I
. i
CAMBRIAN TO EOCENE o

i
,
I

t
EOCAMBRIAN
I

I"",' PRECAMBRIAN BASEMENT 0


3.4 km I 9 kmyolasse
10 km
V.E. 1:1
Molasse
/ I\/1 \--N/TrIF\--117-(1'/\13-1-1
v_i/N/0/11\A-- -1/:1! ZI/ Ir r /
r1C,)/1/\\,(- -I _ii/
\A
/ 7\-/ /
\---1/N

I //i - 1\

7-,
1

kI I_ I
/ \_\\

Figure A5
114

thrust faults; and (3) crustal shortening via thickening of the crust
and escape block tectonics along strike-slip faults to the north of the
collision zone (Molnar and Tapponnier, 1975; Tapponnier and Molnar,
1977).
Balanced sections through the central Salt Range suggest a minimum
of 19 to 23 km of horizontal shortening of the competent Cambrian
through Eocene section (e.g. fig. A5). The calculated shortening is a
minimum because erosion from the Salt Range front has not been
accounted for. The apparent shortening is less across reentrants where
erosion and deposition are greater (Baker, in prep). Therefore, the
true amount of shortening across the frontal ramp is thought to be
closer to 23 km. Additional factors such as a reinterpretation of the
orientations of high-angle faults (see fig. A5) only have a minor
effect on the measured amount of shortening.
Allowing for 1 km of error, 18-24 km is used as the range of
shortening values. Assuming the thrust sheet overrode the ramp between
2.1 and 1.6 mya, the minimum rate of shortening between the Indian
basement and the Phanerozoic sedimentary cover is between 9 and 14
mm/yr. These rates are comparable to the 10-15 mm/yr convergence
absorbed by underthrusting at the Himalaya as estimated by Lyon-Caen
and Molnar (1985) for northern India. Our balanced cross-section
extended to the north indicates 20% shortening for a 90 km length of
the fold-thrust belt between the Lilla structure and intense
deformation in the northern Potwar Plataeu. This is a relatively
low degree of shortening compared to values obtained in other
fold-thrust belts around the world, probably because the thrust front
only recently jumped the large distance from the northern Potwar
Plateau to the Salt Range.
The calculated shortening rate represents about 20-35% of the
convergence rate between the Indian and Asian plates. Therefore, most
convergence is absorbed north of the frontal thrust zone by contraction
within the collision zone and escape-block tectonics. The calculated
rate and percentage of shortening should be compared to the Indian
portion of the Himalayan foreland fold-thrust belt, where a salt
detachment zone is lacking. A comparison of results could be used to
evaluate the decoupling effects of salt and the origin of the
Hazara-Kashmir Syntaxis.

CONCLUSIONS

Surface geology, reflection, and drilling data provide insight into


the structure, timing, and mode of evolution of the Salt Range. A
pre-existing, down-to-the north basement fault controlled ramping of
the encroaching thrust front. The buttressing effect of the basement
offset apparently caused a local buildup of salt; the Salt Range thrust
sheet subsequently overrode the ramp as competent strata were passively
folded into the present anticlinal form. The thrust sheet north of the
ramp is only mildly deformed due to the decoupling properties of salt.

Changing sediment accumulation rates used in conjunction with


stratigraphic evidence suggest that Salt Range ramping occurred about 2
mya. Balanced cross-sections suggest that a minimum of 19 to 23 km of
shortening occurred across the ramp at a shortening rate of 9-14 mm/yr.
115

Therefore, frontal thrusting in the Pakistan foreland accounts for


part of but not the majority of the total plate convergence.

ACKNOWLEDGEMENTS

The Oregon State University project in northern Pakistan is


supported by NSF grants INT-8118403, EAR-8318194, INT-8609914, and
EAR-8608224. NSF grants INT-8517353 and EAR-8407052 support ongoing
studies by Dartmouth College and Peshawar University. Additional
support for the construction of balanced cross sections was provided by
Chevron International, CONOCO Inc. and Amoco Production Company. We
are grateful to the Oil and Gas Development Corporation and the
Ministry of Petroleum and Natural Resources of Pakistan for the use of
reflection and gravity data discussed in this study. The work could
not have been undertaken without scientific collaboration from
individuals within O.G.D.C., the Geological Survey of Pakistan, and the
Centre of Excellence in Geology, University of Peshawar. Special
appreciation is extended to R. A. K. Tahirkheli.
116

APPENDIX B

RESTORATION RESULTS AND IMPLICATIONS

Paleobasement dips (i.e. flexure), the original geometry of salt


deposits, timing of deformational events, and the age of molasse north
of the Salt Range were all used as constraints on restored sections.
The great amount of salt deposited north of the basement offset (com-
pared to relatively minor amounts to the south), in combination with
the independent balancing assumption (see methods), helped to reduce
the retrodeformation of salt to two models. Here, "salt" is used to
represent all strata above the basement and beneath prominent platform
reflectors. The first model, deemed the tabular salt model (fig. 21a)
consists of redistributing salt into two rectangular shaped blocks,
south and north of the basement offset (fig. 13). By summing the salt
area and then dividing by the length over which it must be redistri-
buted, the thickness of the salt blocks were found to be about 0.3 and
1.3 kms, respectively. In the model, the assumption of a discontinuity
in thickness across the basement offset implies Eocambrian basement
faulting (fig. 21a).
The second model represents the retrodeformation of salt to a
wedge shape (fig. 13). The shape of the wedge was found by assuming no
change in thickness across the basement offset, and by using the thick-
ness of autochthonous salt south of foreland deformation (i.e. south of
the Lilla anticline, fig. 5) as a starting point. Knowing this thick-
ness, the length of the wedge, and the total salt area, a value of
constant taper was obtained. This taper was found to be 0.20°, result-
ing in an initial salt thickness of 0.33 km at the basement offset.
Similarly, using 0.33 km as the thickness of salt at the south end of
the second wedge, a taper of 0.93° was found. At this taper the salt-
wedge thickens to 3.0 km at the northern end of the restoration. The
second model requires a change in taper but not thickness at the base-
ment offset. The restored basement dip also changes at this point.
Because Paleogene strata across the Potwar Plateau are primarily
shallow marine limestone and shale the top surface at the end of Eocene
deposition is assumed to be horizontal. The absence of Paleogene
strata south of the basement offset is assumed to be due to post depo-
sitional erosion (fig. 22b), because widespread Paleogene deposition is
indicated (Khan et al., 1986). This assumption, combined with salt
redistribution models, resulted in an increase in basement dip north of
the basement offset at the end of Eocene (fig. 13). Basement dip is
therefore not entirely due to Neogene flexure during Himalayan orogen-
esis, a consideration that is consistent with a possible miogeoclinal
setting of the northwestern Indian margin during the Paleozoic and
Mesozoic.
The assumption of a southward topographic slope toward the basin
depocenter during molasse deposition was derived by comparison to
present day Himalayan foreland basin topography. South of the foreland
basin axis, a reversal in depositional slope is expected where rivers
drain northward from the Indian shield. Because the thickness of
Siwaliks for restoration are fairly well known from seismic and well
control, the requirement of southward topographic slope supplies a
117

valuable constraint on the degree of post 4.5 m.y. flexure. To attain


a southward topographic slope upon restoration of the tabular salt
model, an essentially unflexed basement was required, similar to the
basement structure which existed at the end of Eocene time. In addi-
tion, a substantial amount of unpreserved Upper Siwaliks in the Potwar
Plateau would be required to compensate for potential southward salt
flowage. For the salt-wedge model (fig. 13), the proper direction of
topographic slope is attained unless the basement is completely flexed
into its present shape. Potential southward salt flowage could be
accommodated by unflexing the basement by the necessary amount to
retain the propper direction of surface slope.
A special problem existed in attempting to restore molasse south
of the basement offset because the age of these sediments are unknown.
However, a valuable constraint was supplied because northward transport
of conglomerate clasts into the Baun syncline is interpreted at about
the same time as the 4.5 mya restoration. As a result, the restora-
tions, after northward retrodeformation of the Baun syncline, required
1.7 km of molasse above the up-thrown basement block for this area to
be topographically higher than the Baun syncline (fig. 13).
Two restoration models (i.e. tabular salt and salt-wedge models)
encompassing: (1) the age of basement faulting, (2) redistribution of
salt, (3) timing of flexure, and (4) ramping style, have been consider-
ed. In the tabular salt model, Upper Siwalik conglomerate deposition
at the Baun site would have to be from an uplifted footwall source (see
fig. 22a; Burbank and Raynolds, in press), because the basement is not
yet flexed, and therefore autochthonous Paleozoics should not be expos-
ed. The wedge shaped salt model requires southward directed salt flow
prior to ramping, in order to emplace the thick salt section now
located beneath the Salt Range. Unlike the previous model a plane of
weakness in platform strata would be available for reactivation during
salt growth and subsequent ramping. Finally, instead of an alloch-
thonous source area, a source terrane of Tobra Formation clasts along
the flexural bulge is reasonable in accordance with basement flexure
prior to 4.5 mya.
118

APPENDIX C

FACTORS INFLUENCING SHORTENING CALCULATIONS

In the central Salt Range, erosion is thought to be the biggest


cause of varying shortening values. Rivers debauch from mountain
fronts at reentrants, increasing the potential for erosion. Smaller
shortening values were obtained where transects cross the range front
within the 30 km wide re-entrant in the central Salt Range (figs. 5 and
16). A greater amount of range front retreat within this reentrant
would explain differential shortening values. The smallest amount of
shortening was measured along one of the easternmost sections which
crosses the range front at a gorge where erosion is concentrated via
funneling of runoff and salt diapirism. Since shortening at the toe of
the central portion of the Salt Range thrust sheet increases eastwards,
erosion resulting in an equal distance of range front retreat would
remove strata representing a greater amount of shortening compared to
areas to the west (fig. 17).
The central Salt Range frontal thrust is not exposed except near
Kattha Masral, due to burial by recent sediments. Burial is exempli-
fied by topographic profiles across the range front (figs. 13 and 17),
showing wedges of fanglomerates emanating from the Salt Range. Rivers
discharge their loads at reentrants since they represent the shortest
dispersal path between the hinterland and the foreland basin (Eisbacher
et al., 1974). The effects of burial parallel those of erosion. Maxi-
mum burial should be at reentrants and gorges, where sediments are
debouched. Neglecting burial also minimizes shortening values.
Cross-sections crossing tear faults may not balance (Woodward and
Boyer, 1985). Balancing assumes that reversing fault displacements
will fit pieces back together that were once adjacent. However,
strike-slip displacements move pieces in and out of the plane of sec-
tion. The problem occurs where zones of different shortening accommo-
dation are juxtaposed by strike-slip movements. Whether shortening is
over or underestimated depends on the original distribution of internal
shortening. In part, these factors are minimized since Salt Range
strike-slip faults apparently developed prior to or during ramping and
associated internal deformation (fig. 19; e.g. Dubey, 1979). Close
inspection of figure 16 reveals that potential juxtaposition of differ-
ent regimes of shortening should not significantly effect shortening
values.
A potentially significant factor is the assumption of matching
footwall and hanging-wall ramps. Slight hanging-wall ramp steepening
was used (fig. 13 and 17) to bring sections into balance. Ramp shapes
may not match due to folding prior to or during ramping, resulting in a
smooth rather than a ramp-flat fault trajectory (Cooper and Trayner,
1986). The interpreted ramping style, which utilizes a plane of weak-
ness, suggests that ramps should approximately match, unlike fault-
propagation folds in the eastern SR/PP. Restoration of the hanging-
wall ramp could have a large effect on shortening. The position of the
ramp cutoff was selected so as to minimize shortening values. The
assumption of approximately matching ramp shapes together with line
length balancing helped to constrain the shapes of both hanging-wall
119

and footwall ramps.


Folds cannot simply be unrolled above areas of exposed ductile
flow; overlying beds may have undergone elongation (Ramsay, 1969).
Therefore, unpreserved extensional features above salt uplifts would
result in an overestimation of shortening. This is not thought to be a
major factor since cross-section traces do not cross wide outcrops of
salt and the largest shortening values were obtained along transects
which crossed little salt (fig. 17).
The orientations of high angle faults are generally not discern-
able on seismic lines. There are two major considerations when inter-
preting fault orientations during balancing. First, a predominance of
normal faults decreases measured shortening while reverse faults would
increase calculated shortening. This influence decreases with increas-
ing fault dip. To minimize shortening most of the high angle faults on
the frontal balanced sections were interpreted as normal faults (fig.
17). In addition, fault orientation control on shortening was mini-
mized by maximizing fault dips within reason. The second factor con-
cerns the cross-sectional profile of faults. Fault plane changes are a
major tool for bed length adjustments. Reorientation of straight line
fault plane profiles does not change relative bed lengths. However,
curved fault plane profiles (e.g. listric normal faults) do change
relative bed lengths. For example, a downward concave reverse fault
lengthens upper relative to lower beds while an upward concave reverse
fault has the opposite effect. This principle was used in balancing
frontal sections (fig. 17c). In this way, bed length balancing
can constrain the orientations of major faults.
Failure to restore clipped footwall strata will throw a section
out of balance and decrease measured shortening. For this reason,
horses of clipped footwall strata are tacked on to the range front,
even though they may lie beneath the Salt Range (fig. 17).
120

APPENDIX D

PETROLEUM GEOLOGY

Petroleum exploration in the Pakistan foreland is in its first


phase; only large structural traps have been tested (see Khan et al.,
1986). Statistics (Baker et al., 1984) suggest that great reserves
remain in untested structural and stratigraphic traps. For example,
truncated Paleozoic and Mesozoic strata beneath the pre-Tertiary
unconformity across the Potwar Plateau represent potential strati-
graphic traps. The recent discovery of the Dhurnal field beneath an
upper-level detachment, and its structural similarity to the Canadian
foothills indicate that large reserves may remain undiscovered in the
NPDZ. The lack of success farther south (e.g. Lilla well; fig. 3)
points to the problem of sufficient petroleum migration into young
structural traps. For these reasons, the timing of deformation
throughout the foreland as well as the structural style in the NPDZ
are important priorities.

Potrebbero piacerti anche