Sir Arsalan Radio

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

RADIOACTIVE DATING METHODS FOR

CENOZOIC ROCKS OF HIMALAYA

MUSFIRA BIBI

Submitted to Sir Arsalan


Department of Geology

University of Karachi

Karachi Pakistan

2020

1
2
Table of Contents

Chapter 1: Introduction ................................................................................................................................. 6


1.1. Background ................................................................................................................................. 6
1.2. Overview of the Himalaya .......................................................................................................... 8
1.3. Present-day geology and structures of the Himalaya .............................................................. 9
1.4. Shortening Processes in the Himalaya .................................................................................... 13
1.4.1. Convergence Accommodation in the Mid-Crust ................................................................ 13
Chapter 2: Geological setting of Himalaya in Pakistan .............................................................................. 16
Chapter 3: Stratigraphy of the Tertiary foreland basin ............................................................................... 17
3.1. Hangu Formation ...................................................................................................................... 18
3.2. Lockhart Formation ................................................................................................................. 19
3.3. Patala Formation ...................................................................................................................... 19
3.4. Margalla Hill Limestone and Chorgalli Formation ............................................................... 20
3.5. Kuldana Formation .................................................................................................................. 20
Chapter 4: U–Pb zircon geochronology ...................................................................................................... 21
4.1. Methods.......................................................................................................................................... 25
4.2. Zircon ages of tuffite from Kuldana Formation..................................................................... 25
4.3. Detrital zircon ages ................................................................................................................... 25
4.3.1. Balakot section................................................................................................................... 25
4.3.2. Muzaffarabad section ....................................................................................................... 27
4.3.3. Quantitative analysis......................................................................................................... 27
Chapter 5 : Discussion ................................................................................................................................ 28
5.1. Detrital zircon provenance of Cenozoic sequences ................................................................ 28
References ................................................................................................................................................... 30

3
Table of FIgures

Figure 1 Map of the Himalaya showing the regions referred to in the tex……………………………...…8
Figure 2 Generalized tectonic map of the Himalaya showing major tectonostratigraphic zones. .............. 10
Figure 3 Cross-section of the Himalaya detailing major structural features,. ............................................ 12
Figure 4 Model for the development and exhumation of the Greater Himalayan Sequence. ..................... 14
Figure 5 Schematic representation of duplex models proposed .................................................................. 15
Figure 6 A) Revised detailed geological map of the Balakot section, showing structural geology ............ 18
Figure 7 Lithostratigraphic chart of the Balakot section showing sample locations. ................................. 22
Figure 8 The lithostratigraphic column of the Muzaffarabad section showing stratigraphic position ........ 23
Figure 9 A) Comparison of probability density diagrams for detrital zircon ages. .................................... 24

4
Abstract

The northernmost exposures of sub-Himalayan Cenozoic strata in the Hazara–Kashmir syntaxial


region of north Pakistan comprises the Paleocene–Eocene marine strata in the lower part and
Oligocene– Miocene nonmarine strata in the upper part. This study provides the detrital zircon
U–Pb geochronology of the Cenozoic strata in this area. The strong resemblance of U–Pb age
spectra of Paleocene Hangu, Lockhart and Patala formations with those of Himalayan strata
indicate an Indian plate provenance. The first appearance of <100 Ma detrital zircon U–Pb ages
within the lower most part of the Early Eocene Margalla Hill Limestone indicates a shift from an
Indian to Asian provenance. Geologic mapping shows the existence of a disconformity between
the lower and upper most part of the Patala Formation, which is interpreted to have been formed
by the migration of a flexural forebulge through this region. We consider the upper most part of
the Patala Formation to have been deposited within the distal foredeep of the foreland basin. The
Indian to Asian provenance shift and the presence of a possible foreland basin forebulge provide
strong evidence that India–Asia collision was underway in northern Pakistan at ca. 56–55 Ma.

5
Chapter 1: Introduction
1.1. Background

The Himalayan mountain chain is classical example of Continental Collision Tectonics and links
up the presentday geodynamic processes with those of Late Mesozoic and Cenozoic. The geodynamic
evolution of the Himalaya has at least 5 (five) different chronological stages:

(a) Late Mesozoic subduction and accretion


(b) Cenozoic Collision
(c) Late Collision extensional tectonics
(d) Post-Collision sedimentation in foredeeps
(e) Presentday geodynamics (Jain et al., 2002)

This geodynamic evolution has been modelled into a global framework of underthrusting Indian
Plate beneath the Eurasian Plate along two major suture zones i.e., the Indus-Tsangpo Suture Zone (Main
Mantle Thrust) and the Shyok Suture Zone (Northern Suture Zone). The overthrust Eurasian Plate reveals
its character in a Late Mesozoic Andean-type Kohistan-Ladakh Batholith Complex, which is separated
from the Karakoram metamorphic and its batholithic complexes by the Shyok Suture Zone. The suture
between the two plates is exposed into two suture zones of the Indus-Tsangpo and Shyok; both are
characterised by island-arc setting, fore-arc and back-arc sediments and ophiolite emplacement.

Early Mesozoic reconstruction reveals the existence of the Tethys Ocean of some 2500 km width
between the Indian Plate and the Tibet block. The latter represents a passive continental margin through
the Mesozoic and Early Eocene, as represented by the Tethys Himalayan sequence from Zanskar, Spiti,
Kumaun, Nepal and southern Tibet. The shelf and the slope facies are recognised in the Mesozoic of the
Tethys Himalaya along the passive Indian margin in Zanskar and southern Tibet.

The Karakoram Supergroup and the Lhasa block sequence of Late Carboniferous to late Cretaceous
age, located north of the ITSZ, represent deposits on the southern margin of the Tibet block of the Eurasia
Plate. A stable carbonate platform recognised in the Mesozoic of the Tibet block existed until initial
closure of this ocean. The Tethys oceanic lithosphere has disappeared along the ITSZ by northward
subduction of the Indian oceanic plate beneath the Tibet block and ophiolite emplacement. The ophiolite
complexes are located along the entire belt of Indus- Tsangpo Suture Zone (ITSZ) from Ladakh to
southern Tibet and within the Shyok Suture Zone. These occur in two main tectonic settings. Thrust
Sheets of ophiolite are obducted southward on to the Tethyan Sedimentary Zone of northern continental
margin of Indian Plate. Secondly, discontinuous tectonic lenses occur within the ITSZ. The Spongtong

6
ophiolite klippe in Zanskar is one example of this obduction, while the Nidar and Shergol melanges in
Ladakh and the Shigaze ophiolites in southern Tibet are described from within the IndusTsangpo suture
zone. These ophiolite complexes represent the dismembered Jurassic-Cretaceous oceanic lithosphere of
the Tethys Ocean that was closed along the suture zones. Through 125 to 45 Ma during Jurassic-
CretaceousEocene, accretion and subduction-related crustal processes are documented in geological and
isotopic signatures of the Indus flysch deposits, Dras Volcanics, Ladakh Batholith complex, dismembered
oceanic crust of the Tethyan Ocean and blueschist metamorphism along the 500 km long Indus Suture
Zone in Ladakh (Jain et al., 2002 and references therein).

The Trans-Himalayan plutons, referred to as the Ladakh, Kohistan and Karakoram batholiths and
Gangdese magmatic belt are manifestations of the subducting Indian Plate (Coward et al., 1982;
Honegger et al., 1982; Dietrich et al., 1983; Khan et al., 1989; Rolland et al., 2000). The Trans-Himalayan
plutons are compatible with I-type Cordillearn batholiths of Peru. Initial Nd, Sr and Pb isotopic
compositions of the Gangdese and Ladakh batholiths indicate that these bodies have predominantly
mantle-derived components. U-Pb ages vary from 105 to 45 Ma ages for Trans-Himalayan plutons. These
ages and geochemistry point to northward subduction of the Tethyan oceanic lithosphere beneath the
Eurasian Plate, formation of magma and minor crustal contamination during rise of these batholiths. The
calc-alkaline Kohistan and Dras island arcs are also product of subduction. They are developed in oceanic
regime and later accreted to the Indian Plate. Closure of the Tethys Ocean is mainly confined along the
Indus-Tsangpo Suture Zone (ITSZ) and the Kohistan Arc.

The northern margin of the Indian Plate in the Higher Himalaya encompasses a basement and its
sedimentary cover of Proterozoic age in the Higher Himalayan Crystalline (HHC). This basement is
overlain by a Paleo-Mesozoic platform sequence of the Tethyan Sedimentary Zone. Subsequent crustal
shortening of the Indian Plate during Continental Collision has considerably deformed the Higher
Himalayan basement metamorphics. This is thrust southward along the Main Central Thrust (MCT) and
its numerous splays over a central part of the dismembered Indian Plate whose Proterozoic-Paleozoic
sedimentary basins in the Lesser Himalaya was covered by two marine transgressions during the Permian
and Eocene (Ganseer, 1964; Valdiya, 1980; Thakur, 1993; Jain et al., 2002; and references therein). The
whole succession is thrust southward over the Cenozoic Foredeep Molasse deposits along the Main
Boundary Thrust (MBT). However, the southern margin of the Himalaya is in contact with a Proterozoic
basement of the Aravalli and its platform in the Vindhyan Basin, which is covered by the Holocene
alluvium of the Indo-Gangetic Plains.

7
1.2. Overview of the Himalaya

The Himalaya and adjacent Karakoram host the tallest mountains on Earth. They are the
physiographic manifestation of the continued collision of India with Eurasia. These and other
related ranges form a broad, ~2500 km long topographic arc along the southern boundary of the
Tibetan plateau (Mascle et. al, 2012) that trends approximately WNW to ESE between syntaxes
at Nanga Parbat in the far west and Namche Barwa in the far east.

Figure 1 Map of the Himalaya showing the regions referred to in the text, the locations of glaciers with moraine ages
assigned to the ‘Little Ice Age’ and sites of palaeoclimate proxy records that span the Common Era.

The Himalayan range is commonly regarded as the type-example of a mountain belt


formed through continent-continent collision. The ongoing collision between India and Eurasia
initiated approximately 50 Mya (eg. Green et al., 2008), culminating in the geology and
structures we see in the field today.

The collision of two continents creates a fundamental space problem. Traditionally,


textbook treatments of collisional orogenesis have drawn a simplified picture of plates colliding
and forming mountains, often with one continental plate sliding neatly beneath another plate (e.g.
Marshak, 2007; Reynolds et al., 2008; Smith and Pun, 2006; Tarbuck et al., 2005). The reality of
this process, which we are only now starting to understand in detail, is much more complicated.

8
A significant amount of the convergence between the two continents may be accommodated
through processes not easily relatable to typical, rigid plate tectonics. Indeed, much recent
research has been conducted in an effort to unravel the details of these processes and how they
may reflect convergence accommodation throughout the collisional history of the orogen (e.g.
Larson and Cottle, 2014, 2010a, 2010b; Carosi et al., 2013; Montomoli et al., 2013; Sakai et al.,
2013; Webb et al., 2013; Streule et al., 2012; Kali et al., 2010).

1.3. Present-day geology and structures of the Himalaya

The convergence that has led to the Himalaya is also reflected in the rise of the Tibetan
Plateau. The continental crust currently beneath southern Tibet has been thickened to
approximately 75-80 km (e.g. Zhang and Klemperer, 2005), which has led to an average
topographic elevation of > 4500 m above sea level. The elevation of much of India to the south,
in contrast, which has a more typical thickness crust at ~35 km, is within a few hundred metres
of sea level. The difference in crustal thickness between India and Tibet forms a natural taper
and a considerable horizontal lithostatic pressure gradient between the orogenic hinterland and
foreland. This gradient and the associated gravitational potential is one of the main forces driving
deformation within the Himalayan system, moving material from higher to lower potential
energy, or from the hinterland towards the foreland, and an equilibrium state.

The main structures along the Himalayan arc have accommodated deformation associated with
the development of the orogenic wedge and the constant creep towards equilibrium (Hodges
2000, Hodges et al., 2001). While there are a number of large-scale structures recognized along
the Himalaya the names and definitions of these structures can vary significantly (e.g. Searle et
al., 2008). For clarity in how they are defined in this thesis, the various structures, and are
discussed below. There are four major north-dipping, orogen-parallel fault systems commonly
mapped along the Himalaya From north to south these include: the South Tibetan detachment
system (STDS), Main Central thrust (MCT), Main Boundary thrust (MBT), and Main Frontal
thrust (MFT). The three thrust systems are

9
Figure 2 Generalized tectonic map of the Himalaya showing major tectonostratigraphic zones. Red rectangle
represents the extent of panel B. The black filled circle with number shows the location of stratigraphic sections
given in panel C. 1 – Sulaiman–Kirther Ranges 2 – Kohat, 3 – Hazara Kashmir Syntaxis (HKS), 4 – Himachal
(India), 5 – Nepal, 6 – Bengal Basin. B) Modified geological map of the Hazara–Kashmir syntaxis and surrounding
area, with references provided in supplementary data. The filled rectangle indicates the study area. C) The
Stratigraphic Correlation chart of Tertiary sequence of Himalayan Foreland Basin (modified after Najman, 2006,
Singh, 2013). Hatched area represents unconformity. (For interpretation of the references to color in this figure
legend, the reader is referred to the web version of this article.)

10
Tibetan Plateau High elevation plateau north of the Himalaya.
North Variably referred to as the orogenic Hinterland. The
crust beneath the plateau is thickened due to the
continued collision of India with the Eurasian plate.
Tethyan Sedimentary Sediments of the pre-existing Tethys ocean between
Sequence (TSS) Tibet and India predating the initial collision of the
Himalaya circa 50 Ma. Now preserved mainly along
southern margin of the Tibetan Plateau.
South Tibetan Detachment System of top-to-the-north sense shear structures
System (STDS) commonly found near the transition from the Tibetan
Plateau to the Himalaya. Juxtaposes the TSS to the
north in its hanging wall against the high-grade
metamorphic rocks of the Greater Himalayan
Sequence to the south in its footwall.
Greater Himalayan Sequence Exhumed, former midcrustal high-grade metamorphic
(GHS) core of the orogen. Bound above and below by the
STDS and MCT (see below) respectively.
Main Central Thrust (MCT) “Base of the large-scale zone of high strain and ductile
deformation, commonly coinciding with the base of
the zone of inverted metamorphic isograds” (Searle et
al., 2008)
Lesser Himalayan Sequence Unmetamorphosed-to-low metamorphic grade rocks,
(LHS) within the footwall of the MCT and the hanging wall
of the MBT
Main Boundary Thrust (MBT) Thrust fault separating the rocks of the LHS in its
hanging wall from the Sub Himalaya / Siwaliks in its
footwall
Sub Himalaya (SH) Also referred to as the Siwaliks, this zone represents
material that was eroded off of the evolving mountain
front, accumulated in the Himalayan foreland basin,
and is now being overridden and incorporated into the
deforming orogenic wedge.
Main Frontal Thrust (MFT) Thrust fault separating the SH to the north in the
hanging wall, and the Indo-Gangetic Plain to the south
in the footwall
Main Himalayan Thrust Sub-surface thrust decollment that is assumed to
(MHT) connect the MFT, MBT, and MCT down towards the
Hinterland
Indo-Gangetic Plain The foreland basin built on the Indian Craton. The
crust is thinner beneath the Indo-Gangetic Plain than
beneath the Tibetan Plateau, creating a lithostatic
South gradient from north to south.

11
Figure 3 Cross-section of the Himalaya detailing major structural features, adapted from Hodges et al., 2001. Half
arrows indicate relative movement directions across faults.interpreted root down structurally toward the north to a
common detachment, the Main Himalayan thrust (MHT).

The STDS comprises a series of top-to-the-north sense ductile stretching faults and low-
angle normal faults that were active between c. 23 Ma and 15 Ma (Kellett et al., 2013; Harris
2007; Godin et al., 2006). The Tethyan sedimentary sequence (TSS) occurs north of the STDS,
in its hanging wall, and represents the supracrustal hinterland of the orogen (eg. Larson et al.,
2010a). The TSS is juxtaposed across the STDS against the subjacent Greater Himalaya
Sequence (GHS). Rocks in the GHS represent the exhumed former midcrustal core of the orogen
and typically record evidence of kyanite to sillimanite grade metamorphism and anatexis (Larson
et al, 2011). The GHS comprises the hanging wall of the MCT, a top-to-the-south sense shear
zone that was at least partially coeval with the movement across the STDS between c. 22 Ma and
12 Ma (Harris, 2007, Godin et al., 2006).

The MCT juxtaposes the GHS over the Lesser Himalayan sequence (LHS; Figure 1.2).
The LHS consists of unmetamorphosed to low metamorphic grade sedimentary and meta-
sedimentary rocks that are intercalated with occasional orthogneiss lenses (Searle et al., 2008).
The LHS occurs in the hanging wall of the MBT, which was actively overthrusting the Sub-
Himalayan (SH) zone, or Siwaliks, in its footwall between c. 10 and 3 Ma (Robert et al, 2013).

The Siwaliks, which consist of the detrital mollass shed off of the growing mountain
range, are carried in the hanging wall of the presently active MFT (Robert et al, 2013). The MFT

12
marks the leading edge of shortening in the Himalaya (Figure 1.2), and separates the SH zone
from the Indo-Gangetic Plain to the south.

1.4. Shortening Processes in the Himalaya

Orogens are continuously evolving as they shorten, extend, uplift, and erode in response to
both local and regional changes in convergence rate, climate, internal rheology, etc. Moreover,
rocks can translate laterally or vertically within the orogen through time in response to the
various structures and processes that may be active. This movement of material through an
orogen means that it may be subject to different processes that dominate at different structural
levels as the orogen evolves. For example, the GHS, which was metamorphosed and deformed at
high temperatures and pressures in the mid-crust, is now found at the surface where it is subject
to low-temperature/surface processes (eg. Wobus et al., 2008).

It is, therefore, important to consider the history of the orogen as a whole when
attempting to analyze what was happening at a particular location or structural level and what it
might reflect in terms of processes. In keeping with this, in order to understand the earlier
midcrustal evolution of the orogen recorded in the GHS we must first work backwards from the
present geometries. Understanding the recent history and its potential effect on previously
imparted geologic characteristics is critical to elucidating the whole evolution of the orogen.

1.4.1. Convergence Accommodation in the Mid-Crust

The structural geometries observed at the surface in the Himalaya today (Figure 1.2) have
largely resulted from convergence accommodation processes affecting the GHS and the LHS
(eg. Larson et al. 2010b). Rocks that comprise the GHS may have accommodated convergence,
at least partially, through lateral midcrustal flow in response to the horizontal lithostatic gradient
between Tibet and India (e.g. Beaumont et al., 2004; Godin et al., 2006). During such a process,
rocks originating within the mid-crust are thought to have weakened significantly as a result of in
situ partial melting. The meltweakened mid-crust was then driven horizontally across the
lithostatic gradient between the thick crust of the Tibetan plateau and the relatively thin crust of
the Indian craton.

13
Thermo-mechanical models of lateral midcrustal flow may also employ focused surface
denudation coupled with a low-viscosity midcrustal ‘channel’ to enhance the gradient and
promote associated exhumation (Beaumont et al., 2001). In these variants, surface erosional
processes effectively act as a ‘release valve’ removing material from the orographic front of the
range, while advecting higher temperature isotherms towards the surface. Research, however,
indicates that the rocks were not being exhumed to the surface at the time the two faults
bounding the GHS were active (DeCelles et al., 2004; Martin et al. 2014).

The extruding mid-crust appears to have cooled as the material moved out from beneath
Tibet and perhaps up a crustal ramp to the south. This cooling affected the rheology of the
midcrustal material, making it less ductile as it approached shallower depths (Larson and Cottle,
2014). As lateral flow of this cooled material ceased, it appears that convergence was taken up
through imbrication of the extruded mid-crust (Figure 1.3; e.g. Carosi et al., 2013; Montomoli et
al., 2013; Larson and Cottle, 2014). The accommodation of convergence in the GHS also appears
to have involved at least some out-of-sequence thrusting (eg. Ambrose, 2014; Regis et al., 2014;
Grujic et al., 2011) further complicating the structural and kinematic history.

Figure 4 Model for the development and exhumation of the Greater Himalayan Sequence. Modified from Larson
and Cottle, 2014. See text for discussion.Convergence Accommodation Processes in the Shallow Foreland

In order for midcrustal characteristics to be observed at the surface today, the mid-crust
must have been exhumed. The exhumation of the GHS appears to have been driven by horizontal
shortening and vertical thickening of the subjacent LHS (McQuarrie 2008, Bollinger et al., 2006)

14
with paired erosion. There are two commonly debated models proposed to describe the
deformation within the LHS (Figure 1.4) including 1) the development of a hinterland-dipping
duplex (e.g. McQuarrie 2008, Robinson et al., 2006) and the development of a foreland-dipping
duplex by underplating along a relatively stationary ramp (e.g. Avouac, 2003; Bollinger et al,
2004, 2006; Herman et al., 2010; Webb et al., 2013).

A hinterland-dipping duplexing develops through in-sequence thrust propagation


towards the foreland within the LHS (Figure 1.4a). In this model, the ramp associated with the
MHT moves towards the foreland as thrusts systematically incorporate new material into the
duplex. A foreland-dipping duplex, in contrast, is developed through underplating, which occurs
when material from the down-going plate is transferred up across the thrust boundary and added
to the hanging wall (Herman et al., 2010). The various slices are progressively moved up and
towards the foreland along a relatively stationary ramp (Figure 1.4b) during which time their
geometry is modified such that they become foreland dipping.

Figure 5 Schematic representation of duplex models proposed to explain deformtion within the Lesser Himalayan
Sequence.

15
Chapter 2: Geological setting of Himalaya in Pakistan
In Pakistan, three major terranes are juxtaposed with each other and which dominate the
geology of the Himalaya (Searle et al., 1999). From north to south, these terrenes consists of the
Karakoram Block to the north of the Main Karakoram Thrust (MKT)/Shyok suture zone, the
Kohistan–Ladakh Arc (KLA), and the Indian Plate to the south of Main Mantle Thrust
(MMT)/Indus Suture Zone. The Karakoram block (KB) consists of a Paleozoic– Mesozoic
sedimentary belt in its northern part, Cretaceous– Miocene Karakoram batholiths rocks in its
central part, and the Karakoram metamorphic complex of late Paleozoic–Cenozoic age affected
by pre- and post-collisional metamorphism and intruded by Eocene–Miocene granitic intrusions
in its southern part (Fraser et al., 2001; Zanchi and Gaetani, 2011).

The KLA is a typical island arc sequence consisting of Cretaceous–Paleocene volcanic


and metasedimentary rocks in the northern part along MKT, Late Cretaceous Kohistan batholith
with younger ∼34–29 Ma leucogranitic intrusions in the central part, and Early Cretaceous
mafic–ultramafic complexes in its southern part north of MMT (Bouilhol et al., 2013; Jagoutz
and Schmidt, 2012; Khan et al., 1993; Petterson, 2010). The collision of the KLA with the KB is
considered to have occurred in Late Cretaceous (∼75 Ma), which is based on the age of Jutal
dykes, which crosscut the collision related fabric (Searle et al., 1999). The age of regional
metamorphism (80–60 Ma) in the Eastern Karakoram is also consistent with late Cretaceous
accretion of the KLA with the KB (Borneman et al., 2015; Heuberger, 2004). Similarly,
Paleomagnetic studies in northern Pakistan also support Cretaceous accretion between the KLA
and KB (Zaman and Bamousa, 2015; Zaman and Torii, 1999).

From north to south, the Indian Plate sequence south of MMT is classified into three
tectono-stratigraphic zones: the Higher/Greater, Lesser and Sub Himalayan zones (Greco, 1991).
Although, the exact position of the Main Central Thrust (MCT) is disputed in Pakistan, we
consider it to be the Batal Fault in Kaghan valley, following Chaudhry and Ghazanfar (1990).
Following this classification, the Higher/Greater Himalayan zone lies to the north of Batal
Fault/MCT and consists of pelitic–psammatic schists and gneisses, meta-carbonates and
amphibolites containing eclogite lenses (Kaneko et al., 2003). These rocks are Archean– Middle
Proterozoic in age and locally intruded by the ∼1850 Ma, ∼500 Ma and ∼47 Ma plutonic rocks
(Argles et al., 2003; DiPietro and Isachsen, 2001). The Lesser Himalayan zone is bounded by the

16
Batal Fault in the north and the Main Boundary Thrust (MBT) in the south (Figs. 1 and 2). The
Lesser Himalayan zone north of MBT, consists of Proterozoic basement and Pre-Himalayan
cover strata of the Indian Plate, which is unconformably overlain by the Cambrian–Triassic
carbonate, graphitic phyllite, argillite and metabasalt of Tethyan affinity (DiPietro and Pogue,
2004; DiPietro et al., 2008).

The Proterozoic basement and cover sequence of the Lesser Himalaya are intruded by
Cambrian–Ordovician granites (Le Fort et al., 1980). The Sub Himalayan zone mainly consists
of the Miocene and younger rocks of the Cenozoic foreland basin with few exposures of the
Cambrian and Paleocene–Eocene strata in the Balakot, Muzaffarabad and Kotli areas (Baig and
Munir, 2007). The folding and doming structures are responsible for the exposure of the
Cambrian and Paleocene–Eocene sequence.

The Balakot and Muzaffarabad sections lie in the footwall of MBT, and are part of the
sub-Himalayan zone in north Pakistan (Figs. 1 and 2). The foreland deposits are mainly exposed
south of MBT along the strike in the Himalayan orogen. Thus, in north Pakistan, these sections
are the northernmost exposures and contain strata recording the oldest transition from marine to
continental deposition.

Chapter 3: Stratigraphy of the Tertiary foreland basin


The Cenozoic sequence exposed in the Balakot and Muzaffarabad areas, consists of
lower Paleocene–lower Miocene marine to continental sedimentary deposits. The Cenozoic
sequence unconformably overlies the Lower Cambrian Abbottabad Formation, which consists
dominantly of dolomite with subordinate limestone, shale, and sandstone, in a major anticlinal
structure .

The Cenozoic sequence is divided, from older to younger, into the Hangu Formation,
Lockhart Formation, Patala Formation, Margalla Hill Limestone, Chorgalli Formation, Kuldana
Formation and Murree Formation (Fig. 2, Fig. S1 and Fig. S2; see supplementary data).

17
Figure 6 A) Revised detailed geological map of the Balakot section, showing structural geology and stratigraphic
units. The locations of rock samples for U–Pb geochronology are shown by red circles. B) Revised geological map
of the Muzaffarabad section, showing major formations and their contact relationships. C) Weighted Mean age plot
of Volcanic ash sample of Kuldana Formation. D) Field photograph of the Volcanic ash layer in Muzaffarabad
Section. (For interpretation of the references to color in this figure legend, the reader is referred to the web version
of this article.)

3.1. Hangu Formation

The Hangu Formation consists of a laterite horizon, bauxite nodules, arenaceous


limestone, quartzite, claystone, carbonaceous shale, and coal seams. The shales are gray and
sandstone is light gray and reddish brown on fresh surfaces while rusty brown when weathered.
The sandstone is fine to coarse grained, and in some places hematitic (Fig. S1A). In the
Muzaffarabad section, Hangu Formation is comprised of clay, limonitic sandstone and
carbonaceous shales (Fig. S1B).

The limonitic sandstone is very hard and yellow to brownish in color, while the color of
the shales varies from black to brown (Fig. S1B). The thickness of the formation also varies
within the study area ranging from 2 m to 15 m.

18
The lower contact is unconformable with the Lower Cambrian Abbottabad Formation,
while the upper contact with the overlying Lockhart Formation is conformable (Figs. S1A and
S1B). A rich fossil assemblage has been documented in the Hangu Formation at different
exposures in the Kohat–Potwar Plateau and Salt Range, which suggests an Early Paleocene age
(Shah, 2009 and the references therein).

3.2. Lockhart Formation

The Lockhart Formation is comprised of interbedded nodular limestone and shale. The
limestone is thick bedded to massive and gray to dark gray in color (Fig. S1C). The thickness of
the formation is ∼100 m. The massive portion of the limestone exhibits nodules that vary in
length and width from 2–7 cm and 1–4 cm respectively. In the lower portion, the limestone
contains algal lamellas, while in the upper part the nodules become more prominent in limestone
and the proportion of shale increases.

The upper and lower contact of the Lockhart Formation is conformable (Figs. S1B, S1C
and S1D). The age assigned to the Lockhart Formation is Late Paleocene (∼60–57 Ma) based on
detailed biostratigraphy of the Muzaffarabad section and other areas in Pakistan (Baig and
Munir, 2007; Hanif et al., 2014; Shah, 2009 and the references therein).

3.3. Patala Formation

The Patala Formation consists of brown shale and minor sandstone, siltstone and dark
gray limestone interbeds (Figs. S1D, S1E and S1F). The thickness of the formation varies
between ∼80–130 m. The sandstone interbeds are quartz rich and 5–7 cm in thickness. Towards
the top of the section the limestone and shale beds become more prominent. The lower contact is
sharp and conformable with underlying Lockhart Formation (Fig. S1D). An erosional surface,
indicated by a 30–50 cm laterite horizon, was observed ∼1.5 m below the upper contact with the
Margalla Hill limestone in the Muzaffarabad section (Fig. S1D).

A disconformity may also be indicated by a conglomeratic bed ∼2 m below the upper


contact in the Balakot section (Fig. S1F). The laterite horizon and conglomeratic bed may

19
represents the regional unconformity between the Paleocene and Eocene sequences as
recognized along the strike in Himalayan foreland basin (Garzanti et al., 1987). The part of the
Patala Formation above the erosional surface is considered as the uppermost part. The reported
faunal assemblage from the Patala Formation is assigned to SB-5 to SB-6 of Late Paleocene age
(∼57–55 Ma) (Baig and Munir, 2007).

3.4. Margalla Hill Limestone and Chorgalli Formation

The Eocene nodular limestone unit consists of the Margalla Hill Limestone and
Chorgalli Formation. The Margalla Hill Limestone is distinguished by the dominant appearance
of nodular limestone as compared to the Patala Formation. Thin interbeds of arenaceous shale
are present, which separate the bedding of the nodular limestone. The limestone is dark grayish
in color (Fig. S2A).

The size of the nodules varies between 10–20 cm in length and 15–25 cm in width. In the
upper part, a relatively thick brownish shaly horizon is present. The thickness of the Margalla
Hill Limestone and Chorgalli Formation is ∼50 m in Balakot section and Muzaffarabad section.
The limestone is fossiliferous and contains large forams. The age of the Margalla Hill Limestone
and Chorgalli Formation is constrained to Early Eocene age (∼55–53 Ma) by both foraminiferal
assemblage (Baig and Munir, 2007) and an interbedded tuffite of the lower Kuldana Formation
with weighted mean age of 53.2±2.8 Ma (see below) in Muzaffarabad section (Table S1, Figs.
2C and 2D).

3.5. Kuldana Formation

The Kuldana Formation consists of shale, limestone, marl, sandstone and siltstone (Fig.
S2C). The thickness of the formation ranges from 130–160 m. The shales are multicolored, with
red, maroon, purple and olive green color. The shales are argillaceous to calcareous in nature.
The sandstone and siltstone are red, green and gray in color. The thickness of the sandstone beds
varies between 20–200 cm. The limestone and marl beds are grayish in color and contain
abundant large foraminiferas.

The top of the section consists of an oyster-bearing bed that includes broken shells of
nummulite fossils. Sandstone is also present in the form of lenticular beds within shales. Pebbly

20
sandstone beds are present near the upper contact and the sandstone contains reworked fossils at
contact with the Murree Formation (Figs. S2E and S2F).

The upper contact with Murree Formation is considered to be unconformable and is


marked by the cyclic appearance of sandstone and shale. Biostratigraphic studies on the Kuldana
Formation in various parts of the foreland basin in Pakistan and especially in the Balakot,
Muzaffarabad and Murree areas, suggest an Early Middle Eocene age (53–43 Ma) corresponding
to shallow benthic (SB) zones SB-12, SB-13 and SB-14 (Baig and Munir, 2007; Gingerich,
2003). A volcanic ash layer is identified in the lower part, which yielded weighted mean age of
53.2±2.8 Ma (Table S1, Figs. 2C and 2D).

3.6. Murree Formation

In the Balakot section, Murree Formation represents the northernmost exposure of


continental deposits exposed in the subHimalayan foreland basin. It consists of a cyclic sequence
of red, maroon, green and purple shale and mudstone, red, gray to greenish gray siltstone and
sandstone (Fig. S2G) with intraformational conglomerate at several stratigraphic levels. The
sandstone is fine to medium grained and the presence of detrital muscovite gives it a shiny
appearance. In the middle part of the sequence, the sandstone becomes thick bedded and
commonly contains calcite and quartz veins.

Previous studies proposed an Eocene age for the Murree Formation (Bossart and Ottiger,
1989) based on marl band biostratigraphy, which is later revised considering that the marl band
is a structural slice of the underlying formation (Najman et al., 2002). The Oligocene–Miocene
age has been proposed to the Murree Formation relying on the detrital white mica ages reported
from Balakot and Murree areas (Najman, 2006). ∼200 km south of the study area mammal
fossils has been reported from the Fateh Jhang member (lower part of Murree Formation), which
suggest an Early Miocene age (Shah, 2009 and the references therein).

Chapter 4: U–Pb zircon geochronology


We collected a total of 29 sandstone and sandy shale samples from the Paleocene–
Miocene foreland basin sequence exposed in Balakot and Muzaffarabad sections (Figs. 2, 3 and
4). These include one sample from the Hangu Formation, one from the Lockhart Formation, six

21
from the Patala Formation, four from the Margalla Hill Limestone and Chorgalli Formation, nine
from the Kuldana Formation and eight from the Murree Formation.

Figure 7 Lithostratigraphic chart of the Balakot section showing sample locations and U–Pb probability plots for
selected samples. The corresponding biozones are compiled from literature and references are given in
supplementary data. MH and CHF (U) stands for Margala Hill Limestone and Chorgalli Formation
(undifferentiated).

22
Figure 8 The lithostratigraphic column of the Muzaffarabad section showing stratigraphic position of the samples
selected for U–Pb analysis. The biozones and ages are assigned based on biostratigraphic studies (Baig and Munir,
2007; Gingerich, 2003; Serra-Kiel et al., 1998). The U–Pb probability density diagrams represent the sample
analyzed. MH and CHF (U) stands for Margalla Hill Limestone and Chorgalli Formation (undifferentiated). The red
filled circle represent the position of volcanic ash found in the Kuldana Formation. (For interpretation of the
references to color in this figure legend, the reader is referred to the web version of this article.)

23
Figure 9 A) Comparison of probability density diagrams for detrital zircon ages from the Paleocene–Early Miocene
foreland sequence of Balakot and Muzaffarabad areas (this study), Reference age populations from source terranes
including Himalayas, Kohistan–Ladakh Island arc, Karakoram Block and Lhasa Block (see supplementary data). B)
Results of the quantitative analyses for the detrital zircons U–Pb dataset of Balakot and Muzaffarabad Sections to
the source terranes (Indian and Asian source dataset). Fm – Formation and MH&CHF-Margalla Hill Limestone and
Chorgalli Formation.

24
4.1. Methods

The U–Pb geochronology for detrital zircon grains was performed by using the Agilent
7500a Quadruple Inductively Coupled Plasma Mass Spectrometer (ICP-MS) attached with a
New Wave UP 193 nm ArF excimer laser-ablation system at Institute of Tibetan Plateau
Research, Chinese Academy of Sciences.

The detailed analytical procedure (Cai et al., 2011) and the Cathode Luminescence (CL)
images for selected samples were described in supplementary data. The U–Pb ages are presented
by probability density plots using ISOPLOT (Ludwig, 2003) and Kernel density estimate (KDE)
plots (Fig. S4) using DZstats software (Saylor and Sundell, 2016). The detailed zircon results are
provided in Table S1 of supplementary data.

4.2. Zircon ages of tuffite from Kuldana Formation

The 12 concordant ages of the tuffite sample range between 47–61 Ma (Table S1). The
weighted mean age is 53.2 ± 2.8 Ma, which constrain the lower age limit for the Kuldana
Formation and upper age limit for the Chorgalli Formation. The presence of volcanic glass,
albite, illite, chlorite and rutile indicates the volcanic origin. The results of EPMA analyses (Fig.
S3) are given in supplementary Table S2.

4.3. Detrital zircon ages


4.3.1. Balakot section

Sample H-1 was collected from the Hangu Formation, which is the oldest Paleocene unit
in Pakistan (Fig. 2A). The zircon crystals are euhedral. The 98 concordant ages were obtained
from 100 analyses. The U–Pb zircon age spectrum shows age populations between ∼450–600
Ma with a major peak at 500 Ma, ∼720–950 Ma with peaks at 782 Ma and 875 Ma, ∼1600–1760
Ma with a peak at 1700 Ma, a minor peak at ∼2430 Ma, and scattered ages between ∼2550–
3315 Ma (Fig. 3).

Samples BT-1, BT-2, 2015PB1A, 2015PB06, and BTP-4 are siltstones collected from the
lower to upper part of the Patala Formation (Fig. 3). The zircon crystals are rounded to sub-
rounded. The 482 detrital zircon grains yielded 461 concordant ages. The detrital zircon ages in

25
BT-1 and BT-2 are clustered between ∼900–1930 Ma. The largest age population is clustered
between ∼1500–1800 Ma with peaks at ∼1550 Ma and ∼1750 Ma. A smaller age population is
observed at ∼2360–2550 Ma with a mean age at 2500 Ma (Fig. 3).

The samples from the uppermost part of the Patala Formation exhibit a very different age
spectra compared to those of the lower and middle part. Samples 2015PB1A and 2015PB06
yielded age populations between ∼750–950 Ma with a peak age at ∼875 Ma, between ∼500–600
Ma with a peak age at ∼575 Ma, and ∼1800–1900 Ma (Fig. 3). There are also minor age
populations between ∼2000– 2250 Ma and ∼2400–2800 Ma (Fig. 3). The stratigraphically
highest sample BTP-4 shows a dominant age population between ∼100–160 Ma with peak age at
135 Ma. A significant age group is also present between 1200–2100 Ma (Fig. 3).

Samples BTM-1 and BTM-3 are sandy shales from the bottom and middle part of the
Eocene nodular limestone sequence, respectively (Fig. 3). The 134 out of 165 detrital zircons
yielded concordant ages. Major age populations are and there are some scattered ages between
∼2200–2700 Ma. The youngest suite of ages includes peaks at ∼60 Ma, ∼75 Ma, 100 Ma, 115
Ma and 170 Ma (Fig. 3). Samples BTM-2, BT-7, BT-9, BT-14A and BB-8 were collected from
the lower, middle, and upper part of the Kuldana Formation (Fig. 3). Most of the zircon crystals
were rounded and sub rounded. Collectively, the five samples yielded 477 detrital zircons with
concordant ages. The vast majority (∼68%) of the detrital zircon ages are between ∼43–113 Ma
(Fig. 3). The two most significant age populations are between ∼43–60 Ma with a peak at ∼45
Ma, and between ∼70–113 Ma with a peak at ∼82 Ma.

Subordinate age populations are around ∼500 Ma and ∼1000 Ma, and few scattered ages
between ∼2400–3200 Ma (Fig. 3). Three samples (BT-11, BT-13 and BB-5) of fluvial
sandstones were collected from the Murree Formation. Of the 300 detrital zircon grains
analyzed, 290 yielded concordant ages. The most significant age populations are between ∼400–
600 Ma with a major peak at 500 Ma, ∼700–1200 Ma, ∼1500–1800 Ma, and ∼2300–2500 Ma.
The scattered ages are between ∼2700–3400 Ma (Fig. 3). Only 7 detrital zircon grains out of 290
yielded <100 Ma ages.

26
4.3.2. Muzaffarabad section

Sample 2015MY-14 from the bottom of Lockhart Formation yielded 80 concordant ages
out of 100 analyses (Fig. 4). The ages are mainly clustered around ∼1600–2000 Ma, with a peak
age at ∼1800 Ma. Minor age populations are present between ∼900–1100 Ma and 2100–2500
Ma (Fig. 4). Sample 2015MY-11 was collected from the upper part of Patala Formation near the
contact with Margalla Hill Limestone. The 50 detrital grains yielded 46 concordant ages. The
dominant age populations are between ∼785–1400 Ma and ∼1700–2000 Ma (Fig. 4). There is a
minor age population between ∼450–580 Ma.

Three grains yielded ages in the 100–200 Ma range. Samples 2015MY-10 and 2015MY-
1 were collected from the middle and upper part of the composite Margalla Hill Limstone and
Chorgalli Formation, respectively. Collectively, of the 144 detrital zircons analyzed, 118 yielded
concordant ages (Fig. 4). Almost half (∼54%) of the grains yielded ages of

4.3.3. Quantitative analysis

We performed different statistical analyses to compare our data with source terranes, i.e.,
Himalayan source (Indian Affinity) and/or Asian source (Kohistan–Ladakh arc, Karakoram
Block and Lhasa Block). These analyses include cross correlation, likeness, similarity,
Kolmogorov–Smirnov test (K–S test) and Kuiper’s statistic test using DZstats software (Saylor
and Sundell, 2016). The statistical values range between 0–1, which shows weak to perfect
relation respectively.

While the “k” value for Kolmogorov–Smirnov test and Kuiper’s statistic test shows inverse
relation from 0–1 (i.e., strong to weak relation from 0 to 1). The Paleocene Hangu, Lockhart and
Patala formations from both sections show good relation with Indian affinity detritus (Figs. 5B
and S5–S8). The Margalla Hill Limestone and Chorgalli Formation from Balakot section shows
a good relation to both Indian and Asian sources, while in the Muzaffarabad section, these show
a comparatively strong relation to Asian sources (Figs. 5B and S5–S8). The statistical results of
Kuldana Formation from both sections show a strong relation to Asian sources. The quantitative
analyses show that the Murree Formation has a comparatively strong relation to Indian affinity
source (Figs. 5B and S5–S8).

27
Chapter 5 : Discussion
4.4. Detrital zircon provenance of Cenozoic sequences

In order to mark the India–Asia provenance transition, the U–Pb detrital zircon age data
of various terranes were compiled (Fig. 5A; Supplementary text 2). Of which the Higher
Himalaya (HH), Lesser Himalaya (LH) and Tethyan Himalaya (TH) represents the Indian
affinity and considered collectively as Himalayan source, while Kohistan–Ladakh arc (KLA),
Karakoram Block (KB) and Lhasa Block (LB) represents the Asian affinity. We considered the
KLA as a southern margin of Asian Plate accreted to KB during the latest Cretaceous (Searle et
al., 1999). In the following section the detrital provenance is explained with reference to these
terranes. One sample of the Hangu Formation matches with the Himalayan spectrum, which
suggests derivation from the Indian Plate provenance (Fig. 5A).

A single Cretaceous detrital zircon with age 119 ± 8 Ma in the Hangu Formation
probably represents the derivation from volcanic rocks of the Indian craton (DeCelles et al.,
2000). The one sample of the Lockhart Formation from Muzaffarabad section shows a dominant
age population between ∼1600– 1900 Ma with a peak at ∼1800 Ma (Fig. 5A). This age spectrum
also matches with that of Himalayan spectrum indicating Indian Plate provenance. The detrital
zircons from the lower and middle Patala Formation exhibit age cluster between ∼900–1930 Ma
with a major population around ∼1500–1800 Ma (Fig. 5A). In the upper part, sample 2015PB1A
is from the conglomeratic bed, while samples 2015PB06 and BTP-4 were collected from the
sandy unit above the conglomeratic bed and from arenaceous shale at the upper contact
respectively.

The detrital zircon ages from these samples are clustered around ∼675–975 Ma with a
peak at ∼875 Ma (Fig. 5A). A minor age population present between ∼1600–2200 Ma with a
peak at ∼1800 Ma (Fig. 5A). Sample BTP-4 also shows ages between ∼100–160 Ma with peaks
at 120 Ma, 135 Ma and 143 Ma (Fig. 5A). Similarly, sample 2015MY-11 represents the same
stratigraphic level of the Patala Formation in the Muzaffarabad section, as the samples
2015PB06, 2015PB1A, and BTP-4 in the Balakot section. The age spectrum of the 2015MY-11
is broadly similar to the age spectra of samples collected from the upper part of Patala
Formation. However, the younger grains in this sample is less as compared to BTP-4.

28
The possible reason might be the number of zircon analyzed or the drainage direction in
the basin. The single 67 Ma detrital zircon might be derived from the Deccan Trap as suggested
by the Garzanti and Hu (2014). This contrasting age pattern from lower–middle–upper part of
Patala Formation points to different provenance, but from the same source i.e., Indian Plate.
However the younger age spectrum (100–160 Ma) is similar to the younger age spectrum of
Indian Plate provenance. The ophiolitic remnants preserved within Indus Suture zone also
portray an age spectra within ∼126–170 Ma (Bouilhol et al., 2013).

This suggests that the younger detritus in the upper part of Patala Formation may
possibly derived from these ophiolitic remnants and the Indian Plate. This provenance change
might be related with the migration of forebulge due to loading of KLA on the northern Indian
margin and/or obduction of Ophiolitic remnants, which caused increased input to the upper part
of Patala Formation from younger Himalayan sources and ophiolitic remnants. The age limit of
the Patala Formation is ca. 57–55 Ma, based on biostratigraphic studies in Hazara and Kohat–
Potwar region (Shah, 2009). This age constraint of the Patala Formation is consistent with being
at stratigraphic position between Thanetian Lockhart Formation and Ypresian Margalla Hill
Limestone (Baig and Munir, 2007; Hanif et al., 2014). Despite the contrasting provenance
between lower and upper parts of Patala Formation, there is no evidence of unambiguous Asian
detritus.

The large scale exhumation of Himalaya during Oligocene–Miocene period. Although


the Paleocene Hangu, Lockhart and Patala formations in Pakistan, and the Paleocene Charchare
and Amile formations in Nepal lesser Himalaya lay atop the Paleozoic sequence of Indian
continental lithosphere, which in turn were sourced from the Indian plate, many (∼30%) < 100
Ma detrital zircons of Asian affinity are present within the Eocene Margalla Hill Limestone,
Chorgalli Formation, Kuldana Formation and Oligocene–Miocene continental Murree Formation
in northern Pakistan, but grains < 100 Ma are very rare in the Eocene Bhainskati Formation, and
Miocene Dumri Formation and even younger Siwalik formations, which are dominated by late
Proterozoic and early Paleozoic grains in Nepal (DeCelles et al., 2004; Najman, 2006).

29
References

Ambrose, T.K., 2014. Ductile extrusion, underplating, and out-of-sequence thrusting within the
Himalayan metamorphic core, Kanchenjunga, Nepal. Masters Thesis, UBC Okanagan.

Avouac, J.P., 2003. Mountain building, erosion and the seismic cycle in the Nepal Himalaya.
Advances in Geophysics, 46, 1-80

Beaumont, C., Jamieson, R.A., Nguyen, M.H., & Lee, B., 2001. Himalayan tectonics explained
by extrusion of a low-viscosity crustal channel coupled to focused surface denudation.
Nature, 414, 738-742, doi:10.1038/4147381.

Beaumont, C., Jamieson, R.A., Nguyen, M.H., Medvedev, S., 2004. Crustal channel flows:
Numerical models with applications to the tectonics of the HimalayanTibetan orogen.
Journal of Geophysical Research, 109, 1-29, doi: 10.1029/2003JB002809.

Bollinger, L., Henry, P., & Avouac, J.P., 2006. Mountain building in the Nepal Himalaya:
Thermal and kinematic model. Earth and Planetary Science Letters, 244, 58-71.

Bollinger, L., Avouac, J.P., Beyssac, O., Catlos, E.J., Harrison, T.M., Grove, M., Goffe, B., and
Sapkota, S., 2004. Thermal structure and exhumation history of the Lesser Himalaya in
Central Nepal, Tectonics, 23, 1-19. doi:10.1029/2003TC001564.

Camacho, A., personal communication, 2014. University of Manitoba, Winnipeg, Manitoba.

Carosi, R., Montomoli, C., Iaccarino, S., & Visonà, D., 2013. Tectono-metamorphic
discontinuities in the Greater Himalayan Sequence and their role in the exhumation of
crystalline units. Geophysical Research Abstracts, 15. doi:10.1029/2008TC002400.

DeCelles, P.G., Gehrels, G.E., Najman, Y., Martin, A.J., Carter, A., & Garzanti, E., 2004.
Detrital geochronology and geochemistry of Cretaceous – Early Miocene stata of Nepal:
implications for timing and diachroneity of initial Himalayan orogenesis. Earth and
Planetary Science Letters, 227, 313-330.

Dodson, M.H., 1973. Closure temperature in cooling geochronological and petrological systems.
Contributions to Mineralogy and Petrology, 40, 259-274.
30
Godin, L., Grujic, D., Law, R.D., Searle, M.P., 2006. Channel flow, ductile extrusion and
exhumation in continental collision zones: An introduction. Geological Society Special
Publication, 268, 1-23.

Green, O.R., Searle, M.P., Corfield, R.I., & Corfield, R.M., 2008. Cretaceous-Tertiary carbonate
Platform Evolution and the age of the India-Asia collision along the Ladakh Himalaya
(Northwest India). Journal of Geology, 116, 331-353.

Grujic, D., Warren, C.J., & Wooden, J.L., 2011. Rapid Synconvergent exhumation of Miocene-
aged lower orogenic crust in the eastern Himalaya. Lithosphere, 3, 346- 366.
doi:10.1130/L154.1.

Harris, N., 2007. Channel flow and the Himalayan-Tibetan orogeny: a critical review. Journal of
the Geological Society, 164, 511-523. doi:10.1144/0016-76492006- 133.

Herman, F., Copeland, P., Avouac, J.P., Bollinger, L., Mahéo, G., Le Fort, P., Rai, S.M., Foster,
D., Pêcher, A., Stüwe, K., & Henry, P., 2010. Exhumation, crustal deformation, and
thermal structure of the Nepal Himalaya derived from the inversion of
thermochronological and thermobarometric data and modeling of the
doi:10.1029/2008JB006126.

Hodges, K.V., 2000. Tectonics of the Himalaya and Southern Tibet from two perspectives.
Geological Society of America, 112, 324-350.

Hodges, K.V., Hurtado, J.M., & Whipple, K.X., 2001. Southward extrusion of Tibetan crust and
its effect on Himalayan tectonics. Tectonics, 20, 799-809.

Kali, E., Leloup, P.H., Arnaud, N., Mahéo, G., Liu, D., Boutonnet, E., Van der Woerd, J., Liu,
X., Liu-Zeng, J., & Li, H., 2010 Exhumation history of the deepest central Himalayan
rocks, Ama Drime range: Key pressure-temperature-deformation-time constraints on
orogenic models. Tectonics, 29, 1-31. doi:10.1029/2009TC002551.

Kellet, D.A., Grujic, D., Coutand, I., Cottle, J., & Mukul, M., 2013. The South Tibetan
detachment system facilitates ultra rapid cooling of granulite-facies rocks in Sikkim
Himalaya. Tectonics, 32, 1-28. doi:10.1002/tect.20014, 2013.

31
Larson, K.P., & Cottle, J.M., 2014. Midcrustal Discontinuities and the Assembly of the
Himalayan mid-crust. Tectonics, 33, 718-740.

Larson, K.P., Godin, L., Davis, W.J., & Davis, D.W., 2010a. Out-of-sequence deformation and
expansion of the Himalayan orogenic wedge: insight from the Changgo culmination,
south central Tibet. Tectonics, 29, 1-30. doi:10.1029/2008TC002393.

Larson, K.P., Godin, L., & Price, R.A. 2010b. Relationships between displacement and distortion
in orogens: Linking the Himalayan foreland and hinterland in central Nepal. Geological
Society of America Bulletin, 122, 1116-1134, doi:10.1130/B30073.1.

Larson, K.P., & Cottle, J.M., 2014. Midcrustal Discontinuities and the Assembly of the
Himalayan mid-crust. Tectonics, 33, 718-740.

Marshak, S., 2007. Essentials of Geology. W.W. Norton and Company, New York, London, 2,
70.

Martin, A.J., Copeland, P., & Benowitz, J.A., 2014. Muscovite 40Ar/39Ar ages help reveal the
Neogene tectonic evolution of the southern Annapurna Range, central Nepal. Geologic
Society, London Special Publications, doi:10.1144/SP412.5.

Mascle, G., Pecher, A., Guillot, S., Rai, S.M., & Gajurel, A.P., 2012. The Himalaya-Tibet
Collision. Nepal Geological Society, 1, 1-21. Kathmandu, Nepal.

McQuarrie, N., Robinson, D., Long, S., Tobgay, T., Grujic, D., Gehrels, G., Ducea, M., 2008.
Preliminary stratigraphic and structural architecture of Bhutan: Implications for the along
strike architecture of the Himalayan system. Earth and Planetary Science Letters, 272,
105-117, doi:10.1016/j.epsl.2008.04.030

Montomoli, C., Iaccarino, S., Carosi, R., Langone, A., & Visonà, D., 2013. Tectonometamorphic
discontinuities within the Greater Himalayan Sequence in Western Nepal (Central
Himalaya): Insights on the exhumation of crystalline rocks. Tectophysics,
doi:10.1016/j.tecto.2013.06.006.

32
Regis, D., Warren, C.J., Young, D., & Roberts, N.M.W., 2014. Tecto-metamorphic evolution of
the Jomolhari massif: Variations in timing of syn-collisional metamorphism across
western Bhutan. Lithosphere, 190-191, 449-466.

Reynolds, S.J, Johnson, J.K., Kelly, M.M., Morin, P.J., Carter, C.M., 2008. Exploring Geology.
McGraw-Hill Education, New York, 1, 63.

Robert, X., van der Beck, P., Braun, J., Perry, C., Dubille, M., & Mugnier, J.L., 2013. Assessing
Quaternary reactivation of the Main Central Thrust zone (central Nepal Himalaya): New
thermochronologic data and numerical modeling. Geological Society of America, 37,
731-734. doi:10.1130/G25736A.1.

Robinson, D.M., DeCelles, P.G., & Copeland, P., 2006. Tectonic evolution of the Himalayan
thrust belt in western Nepal: Implications for channel flow models. Geological Society
of America, 118, 865-885. doi:10.1130/B25911.1

Sakai, H., Iwano, H., Danhara, T., Takigami, Y., Rai, S.M., Upreti, B.N., & Hirata, T., 2013.
Rift-related origin of the Paleoproterozoic Kuncha Formation, and cooling history of the
Kuncha nappe and Taplejung granites, eastern Nepal Lesser Himalaya: a
multichronological approach. Island Arc, 22, 338-360. doi:10.1111/iar.12021.

Searle, M.P., Law, R.D., Godin, L., Larson, K.P, Streule, M.J., Cottle, J.M., Jessup, M.J., 2008.
Defining the Himalayan Main Central Thrust in Nepal. Journal of the Geological Society,
165, 523-534, doi:10.1144/0016-76492007-081.

Smith, G.A., & Pun, A., 2006. How does earth work? Physical Geology and the Process of
Science. Pearson Prentice Hall, New Jersey, 329

Streule, M.J., Carter, A., Searle, M.P., & Cottle, J.M., 2012. Constraints on brittle field
exhumation of the Everest-Makalu section of the Greater Himalayan Sequence:
Implications for models of crustal flow. Tectonics. 31, 1-13. doi:10.1029/2011TC003062.

Tarbuck, E.J., Lutgens, F.K., Tsujita, C.J., 2005. Earth: An Introduction to Physical Geology
(Canadian Edition). Pearson Prentice Hall, Toronto, 535.

Villa, I.M., 1998. Isotopic Closure. Terra Nova, 10, 42-47.

33
Webb, A.A.G., Yu, H., He, D., Larson, K., & Schmitt, A.K. 2013. The Himalayan mountains
and analogous systems were built by underplating. Geological Society of America
Abstracts with programs, 47.

Webb, A.A.G., Yu, H., He, D., Larson, K., & Schmitt, A.K. 2013. The Himalayan mountains
and analogous systems were built by underplating. Geological Society of America
Abstracts with programs, 47.

Wobus, C., Pringle, M., Whipple, K., & Hodges, K., 2008. A Late Miocene acceleration of
exhumation in the Himalayan crystalline core. Earth and Planetary Science Letters, 269,
1-10.

34

You might also like