You are on page 1of 24

Tectonics

RESEARCH ARTICLE Multiproxy Isotopic and Geochemical Analysis of the Siwalik


10.1029/2018TC005200
Sediments in NW India: Implication for the Late Cenozoic
Key Points:
• We present a geochemical, Sr-Nd
Tectonic Evolution of the Himalaya
isotopic, and detrital zircon U-Pb Sanjay Kumar Mandal1 , Dirk Scherler2,3 , Rolf L. Romer2, Jean-Pierre Burg4 ,
geochronological data set of the Late
Miocene to Pleistocene Siwalik Marcel Guillong4 , and Anja M. Schleicher2
sediments from two temporally 1
constrained sections in the Dehradun Now at Indian Institute of Science Education And Research Kolkata, Mohanpur, West Bengal, India, 2GFZ German Research
reentrant of NW Indian Himalaya Centre for Geosciences, Potsdam, Germany, 3Institute of Geological Sciences, Freie Universität Berlin, Berlin, Germany,
• Sediments were derived from 4
Department of Earth Sciences, ETH Zurich, Zurich, Switzerland
relatively invariant sources since the
Late Miocene
• Sediments were supplied by
southward directed transverse rivers
Abstract Provenance analysis of the Sub-Himalayan Late Miocene-Pleistocene foreland basin deposits
analogous to rivers feeding the (Siwaliks) from the Dehradun reentrant area provides a 10-Myr long record of the denudation history and
foreland basin at present tectonic evolution of the northwestern Indian Himalaya. We studied Siwalik sediments exposed along the
Mohand-Rao and Haripur-Khol sections, using detrital zircon U-Pb geochronology, major and trace elements,
Supporting Information: and Sr-Nd isotope geochemistry. Results suggest that the erosion pattern has been relatively stable since the
• Supporting Information S1
• Table S4
Late Miocene with sediments derived from the Tethyan Himalayan (THS), Greater Himalayan (GHS), and
• Table S5 outer- (oLHS) and inner-Lesser Himalayan (iLHS) sequences. Provenance data indicate that erosional
• Table S6 unroofing of the Lesser Himalayan Crystalline sequences (LHCS) initiated around 6 Ma, possibly related to
• Table S7
• Table S8
out-of-sequence movement of the Ramgarh-Munsiari Thrust. Our data also suggest erosional recycling of
• Table S9 older foreland basin deposits into younger Siwaliks since ~5.5 Ma, which may indicate the time of thrust
• Table S10 propagation from the Lesser Himalaya into the foreland basin. While the iLHS has been exposed to erosion
• Table S11
since at least ~10 Ma, the Siwaliks were dominated by materials derived from the GHS and THS sources. We
interpret these results as an indication that tectonic uplift and erosion of the orogenic wedge occurred in
Correspondence to: response to duplexing of the iLHS and concomitant high topography and rock uplift rate in the Greater and
S. K. Mandal, Tethyan Himalaya. Comparing the provenance of the Siwalik sediments with that of the modern Ganga and
sanjaykm@iiserkol.ac.in;
Yamuna river sediments further indicates that deposition during the Late Cenozoic was most likely
sanjay.geology@gmail.com
accomplished by southward flowing transverse Himalayan rivers, analogous to the modern ones.

Citation:
Mandal, S. K., Scherler, D., Romer, R. L.,
Burg, J.-P., Guillong, M., & 1. Introduction
Schleicher, A. M. (2019). Multiproxy
isotopic and geochemical analysis of The Himalayan fold-and-thrust belt formed in response to the ongoing continental collision between the
the Siwalik sediments in NW India: Indian and Asian landmasses, which began ca. 60–50 Ma ago (e.g., Hu et al., 2016; Najman et al., 2017).
Implication for the Late Cenozoic
Since the seminal work of Gansser (1964), many studies have expanded our understanding of the geometry,
tectonic evolution of the Himalaya.
Tectonics, 38, 120–143. https://doi.org/ kinematics, and dynamic evolution of this orogen (e.g., Avouac, 2003; Kohn, 2014; Mercier et al., 2017;
10.1029/2018TC005200 Robinson et al., 2003; Webb, 2013; Yin, 2006). This has in part been inspired by the recognition that the global
climatic evolution during the Cenozoic, and related geochemical, oceanographical, and paleobiological
Received 21 JUN 2018
changes, may be linked to the growth of the Himalaya (e.g., Derry & France-Lanord, 1996; Métivier et al.,
Accepted 27 NOV 2018
Accepted article online 30 NOV 2018 1999; Molnar et al., 2010; Quade & Cerling, 1995). Most existing kinematic models have attributed the first-
Published online 11 JAN 2019 order orogenic evolution of this mountain belt to several north dipping, top-to-the-south crustal-scale thrust
zones. These thrust zones have juxtaposed units of different lithology and metamorphic grades (Figure 1).
Subordinate out-of-sequence thrusting (see review in Mukherjee, 2015) has contributed to the general hin-
terlandward to forelandward (in-sequence) propagation of the orogenic growth. Despite variations in tec-
tonic style along and across the orogen (e.g., Yin, 2006, and references therein), structural reconstructions
suggest that three major kinematic phases characterize the orogenic evolution (e.g., DeCelles et al., 2016;
Robinson & McQuarrie, 2012; Webb, 2013, and references therein). The first phase began with the
India-Asia collision and mainly comprised folding and thrusting of the Tethyan sedimentary sequence over
the Indian continental lithosphere (e.g., Ratschbacher et al., 1994; Webb et al., 2011). During the second,
Late Oligocene-Early Miocene phase, metamorphosed Greater Himalayan rocks were transported southward
©2018. American Geophysical Union.
along the Main Central Thrust (MCT, Figure 1; e.g., Catlos et al., 2001; Kohn, 2014; Le Fort, 1986; Vannay et al.,
All Rights Reserved. 2004). During the still ongoing third phase, further shortening produced the Lesser Himalayan duplex (LHD)

MANDAL ET AL. 120


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

Figure 1. Geologic overview map of the Himalaya (location in inset) showing the main tectonic units. Rectangle: Location of Figure 2. STD = South Tibetan
Detachment; MCT = Main Central Thrust; MBT = Main Boundary Thrust; MFT = Main Frontal Thrust; KR = Kangra reentrant; SS = Subathu salient; DR = Dehradun
reentrant. Map compiled from Célérier, Harrison, Webb, et al. (2009); Webb et al. (2011); Hirschmiller et al. (2014).

and the Sub-Himalayan fold-and-thrust belts (e.g., DeCelles et al., 2001; Robinson et al., 2006; Robinson &
Martin, 2014; Srivastava & Mitra, 1994; Webb, 2013). The growth of the LHD, in particular, is thought to have
played a prominent role in the Late Cenozoic phase of Himalayan mountain building (e.g., Avouac, 2003).
Although the timing of LHD formation is critical for shortening estimates across the Himalaya and for
understanding the overall evolution of this mountain belt, so far, it has remained rather poorly documented.
This may in part be related to the dominant clastic and carbonate lithologies of the Lesser Himalayan
unit, which are less suitable for low-temperature thermochronology (e.g., Bollinger et al., 2004; Colleps
et al., 2018).
Foreland basins chronicle the erosional history of their sediment sources, thus offering invaluable informa-
tion about orogenic unroofing histories and inferred deformation events (Allen & Allen, 2013). However,
the composition of sediments that are transported out of an orogen and deposited into the foreland basin
depends not only on the patterns of deformation and subsequent erosion in the orogenic hinterland but also
on the drainage network of rivers that are connecting the sources to sinks (e.g., Allen, 2017).
Provenance analysis of well-dated Sub-Himalayan foreland basin successions affords insights into the onset of
successive erosional unroofing of the main Himalayan tectonostratigraphic units, which in turn provides indirect
age brackets on the orogenic deformation (see review in Clift, 2017). In the northwestern Indian Himalaya,
several researchers used detrital zircon U-Pb geochronology and strontium-neodymium (Sr-Nd) isotope
geochemistry to study the provenance of Eocene to Early Miocene foreland basin deposits (Najman et al.,
2000; Ravikant et al., 2011; White et al., 2002). However, only a few studies (e.g., Najman et al., 2009, 2010) analyzed
the provenance of Late Miocene to Pleistocene Sub-Himalayan foreland basin deposits to constrain the Late
Cenozoic denudation history and thus the tectonic evolution of this part of the Himalaya. Studies of sandstone
petrography and conglomerate clast composition, along with Nd isotope geochemistry of conglomerate clasts
from the Kangra reentrant (Figure 1), suggest that the inner-Lesser Himalayan Sequence (iLHS) was already
eroding at ~11 Ma (Brozovic & Burbank, 2000; Meigs et al., 1995; Najman et al., 2009). This inferred age for the
onset of erosional unroofing of the iLHS agrees well with Nd isotopic results from the central Himalaya in
Nepal (Huyghe et al., 2001, 2005; Robinson et al., 2001) and with zircon (U-Th)/He thermochronological data from
outer-Lesser Himalayan Sequence (oLHS) and iLHS rocks of the northwestern Indian Himalaya (Colleps et al.,
2018). However, based on detrital zircon U-Pb geochronological and Hf isotopic data from Late Paleocene to
mid-Miocene Sub-Himalayan foreland basin deposits in the Subathu salient (Figure 1), Ravikant et al. (2011)
proposed sediment input from the iLHS as early as Late Paleocene/Early Eocene. This age seems at odds with
kinematic reconstructions of the Himalaya, which suggest that exhumation of the iLHS initiated with the

MANDAL ET AL. 121


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

growth of the LHD in the Late Miocene (e.g., DeCelles et al., 2016, and references therein; Srivastava & Mitra, 1994;
Webb, 2013; Webb et al., 2011), which is also suggested by Lesser Himalayan bedrock cooling age data (e.g.,
Colleps et al., 2018).
In this study, we aim to provide further constraints on the timing of erosional unroofing of the iLHS, and thus
LHD formation by studying the Sub-Himalayan Late Miocene-Pleistocene foreland basin deposits (Siwaliks)
exposed along previously magnetostratigraphically dated Mohand-Rao and Haripur-Khol sections
(Sangode et al., 1996, 1999) in the Dehradun reentrant area of the northwestern Indian Himalaya
(Figure 1). We present a systematic sediment provenance study using detrital zircon U-Pb geochronology
and whole-rock Sr-Nd isotope geochemistry. The reliability of the whole-rock isotopic compositions of the
analyzed sediments as source tracers has been evaluated using major and trace element geochemistry.
We used the relatively high-resolution provenance data for reconstructing the denudation history of the
northwestern Indian Himalaya and its tectonic evolution during the past 10 million years (Myr). We further-
more compared provenance data from sand samples from the modern Ganga and Yamuna rivers near the
mountain front (Figure 2a) with data from Siwaliks of the Mohand-Rao and Haripur-Khol sections and recon-
structed the dispersal pathways of the Late Cenozoic foreland basin sediments.

2. Background
2.1. Regional Geological and Tectonic Settings
The studied Mohand-Rao and Haripur-Khol sections are located in the Dehradun reentrant of the northwes-
tern Indian Himalaya, ~14 km southwest and ~20 km west of the cities of Dehradun and Paonta Sahib,
respectively (Figure 2). The thrust belt here strikes northwest-southeast and verges southwestward
(Figure 2a). As elsewhere throughout the Himalaya, the tectonostratigraphy is conventionally divided, from
north to south, into the Tethyan Himalayan, Greater Himalayan, Lesser Himalayan, and Sub-Himalayan
sequences (Figure 2a; Gansser, 1964; Le Fort, 1975; Yin, 2006).
The Tethyan Himalayan Sequence (THS) consists primarily of sedimentary to low-grade metasedimentary
rocks in the hanging wall of the South Tibetan Detachment (STD; Burg & Chen, 1984; Burchfiel et al., 1992).
The Greater Himalayan Sequence (GHS), in the hanging wall of the Main Central Thrust (MCT), consists mostly
of high-grade metamorphic rocks, including ~830- and ~500-Ma orthogneisses and comparatively small
Cenozoic leucogranites and migmatites (Figure 2a; Searle et al., 1999; Spencer et al., 2012a; Vannay et al.,
2004; Webb et al., 2011). The THS strata lie unconformably upon Neoproterozoic-Cambrian low- to
medium-grade metasedimentary rocks (e.g., the Haimanta Group), which are also found in the hanging wall
of the MCT (Figure 2a; Webb et al., 2011). In the footwall of the MCT, the Lesser Himalayan Sequence (LHS)
consists of medium- to high-grade metamorphic rocks (i.e., the Lesser Himalayan Crystalline Sequence;
LHCS) that have been thrust along the Ramgarh-Munsiari Thrust over lower-grade metasedimentary rocks
(Figure 2a; Ahmad et al., 2000; Célérier, Harrison, Webb, et al., 2009; Srivastava & Mitra, 1994; Valdiya,
1980). The Tons Thrust further separates the LHS metasedimentary rocks into the Neoproterozoic-
Cambrian outer-LHS (oLHS) in the hanging wall and the Paleoproterozoic inner-LHS (iLHS) in the footwall,
the latter of which also includes the LHCS (Figure 2a; Ahmad et al., 2000; Célérier, Harrison, Webb, et al.,
2009; McKenzie et al., 2011; Richards et al., 2005; Webb et al., 2011; Valdiya, 1980). Imbricated thrust sheets
of the iLHS form the Lesser Himalayan duplex (Srivastava & Mitra, 1994; Webb, 2013; Webb et al., 2011). In
the vicinity of the Sutlej and Beas rivers, the iLHS rocks crop out within the Kullu window and the
Narkanda half window (Figure 2a). Farther to the southeast, these rocks are exposed in the Uttarkashi half
window (Figure 2a). The oLHS rocks are mostly exposed within structural synforms in northwestern India
(Célérier, Harrison, Webb, et al., 2009; Colleps et al., 2018; Valdiya, 1980). Several GHS klippen are preserved
above the LHS (Figure 2a). The oLHS rocks yield detrital zircon age patterns and isotopic signatures that
are often considered to be equivalent and therefore indistinguishable from the GHS rocks (section 6). As a
result, the tectonostratigraphy of oLHS rocks has been debated (see discussion in DeCelles et al., 2016).
While some researchers suggest that the term (oLHS) should be discarded and that these rocks should be
referred to as a relatively low-grade part of the GHS (DeCelles et al., 2016, p. 3021), we retain the term for this
study, as it is widely used in existing studies of the geology of the northwestern Indian Himalaya. This facil-
itates direct comparison of our provenance data with the published geochronological and isotopic data
and avoids further confusion.

MANDAL ET AL. 122


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License

123
10.1029/2018TC005200
Tectonics

MANDAL ET AL.
19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

The Main Boundary Thrust (MBT) places the oLHS rocks on the Cenozoic foreland basin deposits that
comprise the Sub-Himalayan Sequence (SHS; Figures 2a and 2b). Enlargement of the orogenic wedge
through frontal accretion has incorporated the SHS strata into the fold-and-thrust belts and formed a
hinterland-dipping imbricate zone (Figures 2b and 2c; Avouac, 2003; Hirschmiller et al., 2014; Mishra
& Mukhopadhyay, 2012).
In the Nahan salient area, the Sub-Himalayan imbricate zone exposes the Subathu, Dharamsala, and Siwalik
groups (Figures 2a–2c; Mishra & Mukhopadhyay, 2012; Powers et al., 1998). The Paleogene to mid-Miocene
Subathu and Dharamsala groups are separated from the mid-Miocene to Pleistocene Siwalik Group by the
Bilaspur Thrust (Figure 2b). Finally, the mostly buried but active Main Frontal Thrust (MFT) places middle to
upper Siwalik rocks on undeformed Quaternary sediments of the Indo-Gangetic foreland basin
(Figures 2b–2d; Mishra & Mukhopadhyay, 2002, 2012). In northwestern India, the MFT is thought to absorb
approximately 80% of the 18.5 ± 1.8 mm/a present-day convergence between India and Tibet (Lavé &
Avouac, 2000; Stevens & Avouac, 2016). The initiation of the Sub-Himalayan fold-and-thrust belt is undated,
but the sequence of thrusting has been deduced from balanced cross sections (e.g., Mishra & Mukhopadhyay,
2012; Mukhopadhyay & Mishra, 2005).
2.2. Sedimentology
The studied Mohand-Rao and Haripur-Khol sections are exposed in the central parts of the Mohand and
Dhanaura anticlines, respectively, in the hanging wall of the MFT (Figure 2b). Detailed sedimentological
accounts of these two sections have been provided elsewhere (Kumar, 1993; Kumar et al., 1999; Kumar &
Nanda, 1989), and only summary descriptions are presented below.
2.2.1. Mohand-Rao Section (~9.8–4.8 Ma)
The Siwalik Group rocks exposed along the Mohand-Rao section (Figure 3) document an upward coarsening
clastic succession of fluvial and alluvial fan deposits (Kumar, 1993; Kumar et al., 2004). The lower ~350 m of
the succession consist of 2- to 5-m thick, gray, fine to medium, and occasionally coarse-grained sandstone
bodies and minor intercalations of brown to gray-brown mudstone. The middle ~1,000 m of the succession
consist mainly of thickly (>15 m) bedded, multistoried, friable, salt and pepper textured, massive and cross-
bedded, fine to medium, occasionally coarse-grained sandstones with centimeter to meter scale calcareous
concretionary layers, channel lag deposits of mud balls, mud pellets, and gravel to cobble-sized extraforma-
tional clasts. The upper ~450 m of the succession consist of interbedded gray colored, fine- to coarse-grained
sandstones, siltstones, gray to brown mudstones, thin matrix-supported gravel to cobble conglomerates with
fine- to coarse-grained sand, and silt lenses. The entire succession, ~1,800-m thick in total, has previously
been dated to between ~9.8 and 4.8 Ma by magnetostratigraphy (Sangode et al., 1999). In the absence of
independently dated benchmarks, the age model has been tested against available stratigraphic controls,
iterative matching, comparable sedimentation rates in adjacent sections, and the correlation of magnetic fab-
rics (Sangode et al., 1999). The succession above ~1,800 m comprise thick conglomerates that have not been
dated (Figure 3). The paleocurrent indicators within the sandstone-mudstone dominated interval show high
variability, but mostly southward flow directions (Figure 3).
2.2.2. Haripur-Khol Section (~6.5–<0.7 Ma)
The Siwalik Group rocks exposed along the Haripur-Khol section (Figure 3) also constitutes an upward
coarsening succession of fluvial and alluvial fan deposits (Kumar et al., 1999, 2003). The lower ~500 m
of the succession consist of thickly bedded (20–40 m) multistoried, salt and peeper textured sandstones
(>80%) with lesser amounts of mudstones (<20%). The sandstone bodies contain multiple, vertically
stacked channel deposits separated by internal erosional surfaces. The nearly planar to curvilinear bases
of the sandstone bodies are usually erosive. Lag deposits of calcrete, mud balls, mud pellets, and extra-
basinal quartzite clasts are occasionally found along the basal erosional surfaces. The intervening massive
mudstones are gray to brown and contain calcrete nodules. Immature paleosol horizons are common. The
middle ~1,600 m of the succession is dominated by gray sheet sandstone bodies and mudstone with

Figure 2. Geologic overview of the NW Indian Himalaya. (a) Geological map after Valdiya (1980); Srivastava and Mitra (1994); Vannay et al. (2004); Célérier, Harrison,
Webb, et al. (2009); Webb et al. (2011); Yu et al. (2015). (b) Enlargement of the geology of the Sub-Himalaya in the study area. Thick red lines = location of
magnetostratigraphic sections; A-A’ and B-B0 = location of the balanced structural cross-sections shown in c and d. (c) Balanced structural cross-section along the
Nahan transect (A-A’) after Mishra and Mukhopadhyay (2012). PaT = Paonta Thrust; BiT = Bilaspur Thrust. (d) Balanced structural cross-section along the Dehradun
transect (B-B0 ) after Mishra and Mukhopadhyay (2002). US = Upper Siwalik; MS = Middle Siwalik; LS = Lower Siwalik.

MANDAL ET AL. 124


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License

Figure 3. Correlation of paleomagnetic stratigraphy of Mohand-Rao and Haripur-Khol sections with the revised global

125
10.1029/2018TC005200

magnetic polarity timescale of Hilgen et al. (2012). Magnetostratigraphy adapted from Sangode et al. (1996, 1999).
Tectonics

MANDAL ET AL.
19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

minor buff ribbon sandstone bodies. Thin layers of matrix-supported conglomerates are occasionally
found within the sandstones. The upper ~400 m of the succession begins with conglomerates dominated
by subangular to subrounded sandstone clasts. This interval is composed mainly of interbedded matrix-
to clast-supported conglomerates, fine- to coarse-grained sandstones, siltstones, and lenticular to massive
mudstones. The entire succession, ~2,400-m thick in total, has been dated by magnetostratigraphy to
between ~6.5 and ~0.7 Ma (Sangode et al., 1996; updated using the global geomagnetic reference time-
scale of Hilgen et al. (2012)). A bentonite tuff horizon, dated to ~2.5 Ma, allowed to independently con-
strain the magnetostratigraphic age model near the Gauss/Matuyama boundary at ~1,750 m in the
measured section (Kumar et al., 1999). The massive conglomerates above ~2,400 m have no age con-
straint. Paleocurrent indicators (trough cross-strata and imbricated clasts) point to mainly southward flow
directions, with variations to southwest and southeast (Figure 3).
In this study, we focused only on the magnetostratigraphically dated intervals of both sections, which consti-
tute a ~10-Myr record of synorogenic sedimentation.

3. Data and Methods


3.1. Analytical Approach
We used bulk silicate Sr-Nd isotopic analysis and detrital zircon U-Pb dating on samples collected from both
sections to identify temporal variations in sediment provenance and thus the relative changes of exposed
rock units. Owing to the wealth of existing data on the isotopic composition and zircon U-Pb ages from
the THS, GHS, oLHS, and iLHS rocks (see section 6), both analytical methods have become standard tools
to discriminate source terrains in the Himalaya (e.g., Najman, 2006).
Provenance interpretation based on bulk silicate isotopic data relies on the assumption that the analyzed
sediments retain the isotopic signatures of their source rocks. However, a sandstone sample may not be
representative of the entire sediment load at the time of deposition. Spatial heterogeneities induced by
hydrodynamic sorting, for example, may result in mineralogical and thus chemical differentiation, and in turn,
potentially bias the interpretations (e.g., Garzanti et al., 2010). Therefore, we restricted our sampling to mainly
fine-grained sandstones, which should represent the bulk chemical signature of the source rocks more accu-
rately because they are better mixed and chemically more homogenous than coarser fractions (e.g., Garzanti
et al., 2011). To test this assumption, we also analyzed the samples for bulk chemical composition and com-
pared variations in Sr-Nd isotopic compositions with variations in selected chemical parameters.

3.2. Sampling Strategy


To ensure correct positioning of our samples in the published paleomagnetic stratigraphy of the Mohand-
Rao and Haripur-Khol sections, we reestimated the sediment thickness with a combination of measuring tape
and Bushnell ProX2 laser rangefinder (accuracy: ±0.5 m). Our reestimated sediment thicknesses differ from
the published ones by 60 m (3%) and 15 m (<1%) for the Mohand-Rao and Haripur-Khol sections, respec-
tively. This exercise enabled us to correlate the individual units with their published lithologs (Kumar et al.,
1999; Kumar & Nanda, 1989). We were thus able to locate the magnetic reversal sites in the stratigraphy
reported by Sangode et al. (1996, 1999). This, in turn, enabled us to accurately tie the sampling points to
the magnetostratigraphy and the sample ages were estimated by extrapolating the sediment accumulation
rates recorded by the magnetostratigraphy. Thrity-one samples were collected from the Mohand-Rao section
and 41 samples from the Haripur-Khol section (Figure S1 in the supporting information). To capture the sedi-
ment provenance in high resolution, we limited the stratigraphic spacing of samples to approximately 50 m.
However, we could not avoid some gaps, in places exceeding 100 m, especially in the Mohand-Rao section,
due to lack of or limited access to outcrops, or unfavorable lithology. A summary of location, stratigraphic
position, depositional age, and lithology of the samples is provided in Table S1 in the supporting information.
Furthermore, we sampled two modern river sediments by dredging from the active channel bottom. Our pri-
mary objective was to fingerprint the present-day erosional flux from the Ganga and Yamuna rivers, which
drain the Himalayan hinterland and shed sediments in the vicinity of the two studied sections. By reference
to the known ages of the bedrock sources of each watershed, we attempted to trace their paleosediments in
the stratigraphic record of the two sections. The Ganga River was sampled at Lakshman Jhula bridge
(30.132528°N; 78.330806°E), ~2-km upstream from the mapped location of the MBT (Figure 2a). The

MANDAL ET AL. 126


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

Yamuna River was sampled near the town of Paonta Sahib (30.430000°N; 77.642028°E), ~15-km downstream
from the mapped location of the MBT to include sediments shed by the Giri River (i.e., the southernmost hin-
terland draining tributary of the Yamuna River) into the Yamuna River (Figure 2a).

3.3. Bulk Sample Chemistry and Sr-Nd Isotopic Analysis


We analyzed major and trace element concentrations as well as Sr-Nd isotopic compositions of all sandstone
samples (Figure 3 and Table S1 in the supporting information). The chemical composition was measured at
GFZ German Research Centre for Geosciences using X-ray fluorescence spectrometry and Inductively
Coupled Plasma Mass Spectrometry (ICP-MS; Romer & Hahne, 2010). Details about analytical methods are
provided in the supporting information, with results in Tables S2–S4. The obtained compositions are dis-
cussed with reference to the average concentration of chemical elements in the upper continental crust
(Rudnick & Gao, 2003).
Sr and Nd isotopic compositions of the bulk silicate fraction of the same samples were also measured at GFZ
German Research Centre for Geosciences using thermal ionization mass spectrometry. The analytical proce-
dure is described in the supporting information; results are reported in Table 1. The Nd isotopic compositions
are reported as εNd(0) using the present-day chondritic uniform reservoir value of 0.512638 (Jacobsen &
Wasserburg, 1980).

3.4. Detrital Zircon U-Pb Geochronology


We determined the detrital zircon U-Pb ages of 21 selected samples (10 from the Mohand-Rao section and 11
from the Haripur-Khol section) from the same sample suite and the two modern river sediments (Table S1
and Figure 3). All analyses were performed at the Department of Earth Sciences, ETH Zürich, using laser
ablation-inductively coupled plasma- mass spectrometry (LA-ICP-MS; Guillong et al., 2014). Analytical details
are provided in the supporting information. Results including analytical data of secondary zircon reference
materials (Black et al., 2004; Jackson et al., 2004; Paton et al., 2010; Sláma et al., 2008; Wiedenbeck et al.,
1995) and metadata of the method are reported in the supporting information Tables S5–S9, following the
recommendations in Horstwood et al. (2016). For each sample, at least 120 zircon grains were analyzed.
Grains with age discordance >10% were rejected. Concordia diagrams are shown in the supporting informa-
tion Figures S2–S4. Detrital zircon U-Pb age results were plotted as probability density plots and histograms
with a bin width of 50 Myr for each sample.

4. Results
4.1. Geochemical Characterization of Mohand-Rao and Haripur-Khol Sandstones
The chemical composition of the Mohand-Rao and Haripur-Khol sediments reflects their mineralogical com-
position. The Al/Si and Fe/Si ratios of Himalayan-derived sediments correlate directly with bulk sediment
grain size (e.g., Lupker et al., 2013) and thus constitute a proxy of mineral sorting. Our samples are character-
ized by relatively narrow ranges in Al/Si (0.04 to 0.14) and Fe/Si (0.02 to 0.09) ratios, except MR17-20 and
HK17-30, which are enriched in Fe and Al, respectively (Figure 4a). Our samples fall into two groups in the
Al/Si versus Fe/Si diagram. Group 1 samples, with low Al/Si and Fe/Si ratios, are mainly quartz-rich sediments.
Group 2 samples, with relatively high Al/Si and Fe/Si ratios, are quartz-poor and finer sediments. Both groups
are clustered in the Zr/Sc versus Th/Sc diagram, except for a few samples from group 1, which fall on a trend
toward higher Zr/Sc and Th/Sc values (Figure 4b). The highly correlated trends in the Al/Si versus alkali and
alkaline earth metal diagrams (Figures 5a–5f) indicate that the majority of samples have rather similar
mineralogy. However, some samples deviate from these linear trends (Figure 5a–5f), which indicate variable
admixtures of major minerals such as biotite and alkali feldspars (see supporting information). Note that the
scatter in elements like Ti and Zr (Figures 5g and 5h) as well as contrasting upper continental crust-
normalized rare earth element (REE) patterns (Figure S5 in the supporting information) are related to varying
abundance of accessory minerals. For details see supporting information.

4.2. Sr-Nd Isotope Geochemistry


The εNd(0) values and 87Sr/86Sr ratios in 72 samples range from 23 to 15 and 0.751 to 0.846, respectively,
and the ranges from both sections broadly overlap (Figure 6). Group 1 samples have on average lower
87
Sr/86Sr and higher εNd(0) values than the group 2 samples (Figures 6 and S6). The Sr-Nd isotopic profiles

MANDAL ET AL. 127


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

Table 1
Sr and Nd Concentrations and Isotope Data of the ~9.8- to 0.7-Ma Siwalik Sandstones From the Mohand-Rao and Haripur-Khol Sections
b b b b b
Age Rb Sr Nd Sm TDM
a 87 86 b 147 144 143 144 b
Samples Formation (Ma) (ppm) (ppm) Sr/ Sr 2 SE (ppm) (ppm) Sm/ Nd Nd/ Nd 2 SE εNd(0) (Ga)

Mohand-Rao section
MR17-01 Middle Siwalik 9.81 138.4 35.2 0.776552 0.000015 32.7 6.21 0.115 0.511758 0.000004 17.2 1.93
MR17-02 Middle Siwalik 9.59 62.1 24.5 0.772183 0.000012 25.0 4.68 0.113 0.511741 0.000005 17.5 1.92
MR17-03 Middle Siwalik 9.02 139.3 34.5 0.775659 0.000006 25.6 4.65 0.110 0.511747 0.000002 17.4 1.85
MR17-04 Middle Siwalik 8.75 59.7 45.3 0.779725 0.000010 27.5 5.17 0.114 0.511600 0.000002 20.2 2.12
MR17-05 Middle Siwalik 8.69 63.0 29.3 0.772501 0.000013 18.9 3.36 0.108 0.511711 0.000003 18.1 1.87
MR17-06 Middle Siwalik 8.53 67.1 41.1 0.814032 0.000013 22.9 4.26 0.112 0.511563 0.000002 21.0 2.14
MR17-07 Middle Siwalik 8.38 50.0 32.3 0.766340 0.000014 17.1 3.34 0.118 0.511726 0.000004 17.8 2.03
MR17-08 Middle Siwalik 8.29 58.8 45.6 0.776440 0.000008 26.2 4.99 0.115 0.511608 0.000002 20.1 2.14
MR17-09 Middle Siwalik 8.19 47.5 24.2 0.785977 0.000010 18.0 3.40 0.114 0.511708 0.000003 18.1 1.98
MR17-10 Middle Siwalik 7.88 65.4 35.8 0.778208 0.000011 15.6 3.04 0.118 0.511743 0.000003 17.5 1.99
MR17-11 Middle Siwalik 7.68 71.6 38.2 0.769454 0.000012 22.8 4.35 0.115 0.511729 0.000003 17.7 1.97
MR17-12 Middle Siwalik 7.66 85.5 40.4 0.773161 0.000008 14.6 2.87 0.119 0.511692 0.000003 18.5 2.10
MR17-13 Middle Siwalik 7.59 70.5 40.1 0.757584 0.000012 22.5 4.19 0.112 0.511743 0.000003 17.5 1.90
MR17-14 Middle Siwalik 7.21 74.8 39.0 0.772268 0.000015 22.2 4.17 0.114 0.511639 0.000004 19.5 2.07
MR17-15 Middle Siwalik 7.18 87.3 44.4 0.780854 0.000013 17.2 3.30 0.116 0.511674 0.000002 18.8 2.06
MR17-16 Middle Siwalik 7.17 79.8 40.1 0.845953 0.000006 31.5 5.61 0.108 0.511468 0.000002 22.8 2.19
MR17-17 Middle Siwalik 7.15 52.8 43.8 0.776883 0.000013 20.6 3.93 0.115 0.511673 0.000003 18.8 2.05
MR17-18 Middle Siwalik 6.94 67.6 45.2 0.783349 0.000005 22.5 4.17 0.112 0.511626 0.000002 19.7 2.06
MR17-19 Middle Siwalik 6.66 72.7 45.1 0.786585 0.000008 18.1 3.53 0.118 0.511660 0.000002 19.1 2.11
MR17-20 Middle Siwalik 6.53 36.2 55.9 0.768607 0.000010 38.2 7.44 0.118 0.511648 0.000004 19.3 2.13
MR17-21 Middle Siwalik 6.41 48.9 43.4 0.796284 0.000008 22.3 4.39 0.119 0.511719 0.000002 17.9 2.05
MR17-22 Middle Siwalik 6.21 135.3 53.8 0.797389 0.000013 30.8 5.74 0.113 0.511640 0.000003 19.5 2.05
MR17-23 Middle Siwalik 6.04 79.4 37.5 0.795926 0.000012 15.4 2.91 0.114 0.511713 0.000004 18.0 1.98
MR17-24 Middle Siwalik 5.71 96.6 34.3 0.797959 0.000010 14.0 2.72 0.117 0.511642 0.000003 19.4 2.13
MR17-25 Middle Siwalik 5.47 65.3 34.7 0.786337 0.000008 18.0 3.34 0.112 0.511718 0.000003 17.9 1.94
MR17-26 Middle Siwalik 5.24 55.6 24.9 0.801540 0.000013 17.8 3.44 0.117 0.511680 0.000004 18.7 2.07
MR17-27 Upper Siwalik 5.18 87.4 44.3 0.788365 0.000009 20.8 4.02 0.117 0.511653 0.000004 19.2 2.11
MR17-28 Upper Siwalik 5.13 91.5 48.3 0.824355 0.000014 15.0 2.85 0.114 0.511671 0.000004 18.9 2.04
MR17-29 Upper Siwalik 5.08 51.4 46.1 0.793405 0.000008 23.3 4.18 0.109 0.511557 0.000003 21.1 2.09
MR17-30 Upper Siwalik 5.02 71.4 55.7 0.759560 0.000006 18.3 3.56 0.117 0.511823 0.000003 15.9 1.87
MR17-31 Upper Siwalik 4.97 107.6 52.3 0.782489 0.000010 14.9 3.14 0.127 0.511711 0.000003 18.1 2.24

Haripur-Khol section
HK17-01 Middle Siwalik 6.58 65.6 27.2 0.786029 0.000005 14.8 2.81 0.115 0.511715 0.000005 18.0 1.99
HK17-02 Middle Siwalik 6.42 56.1 34.6 0.775272 0.000005 27.9 5.22 0.113 0.511614 0.000005 20.0 2.09
HK17-03 Middle Siwalik 6.32 101.7 51.2 0.774633 0.000004 34.9 6.06 0.105 0.511606 0.000003 20.1 1.96
HK17-04 Middle Siwalik 6.21 97.2 41.7 0.769270 0.000005 46.3 8.38 0.109 0.511794 0.000002 16.5 1.78
HK17-05 Middle Siwalik 5.98 79.3 41.5 0.784255 0.000005 19.1 3.62 0.114 0.511650 0.000005 19.3 2.07
HK17-06 Middle Siwalik 5.78 52.0 51.6 0.750630 0.000005 23.5 4.38 0.113 0.511658 0.000003 19.1 2.03
HK17-07 Middle Siwalik 5.66 39.4 17.6 0.773514 0.000005 15.9 2.99 0.114 0.511769 0.000003 17.0 1.89
HK17-08 Middle Siwalik 5.50 24.6 10.8 0.762634 0.000006 21.4 3.87 0.110 0.511768 0.000003 17.0 1.82
HK17-09 Middle Siwalik 5.36 26.2 10.2 0.761658 0.000006 9.5 1.73 0.110 0.511847 0.000008 15.4 1.72
HK17-10 Middle Siwalik 5.24 60.0 44.9 0.771940 0.000005 19.3 3.70 0.116 0.511706 0.000004 18.2 2.01
HK17-11 Upper Siwalik 4.94 79.5 41.9 0.783561 0.000004 22.4 4.29 0.116 0.511642 0.000005 19.4 2.10
HK17-12 Upper Siwalik 4.91 81.9 34.2 0.786103 0.000008 15.5 2.98 0.116 0.511738 0.000004 17.6 1.97
HK17-13 Upper Siwalik 4.84 89.7 40.7 0.782326 0.000008 18.8 3.69 0.119 0.511695 0.000006 18.4 2.08
HK17-14 Upper Siwalik 4.68 71.3 37.8 0.802946 0.000008 14.6 2.73 0.113 0.511695 0.000005 18.4 1.99
HK17-15 Upper Siwalik 4.59 25.4 10.0 0.756863 0.000009 12.0 2.22 0.112 0.511828 0.000005 15.8 1.78
HK17-17 Upper Siwalik 4.35 86.0 43.3 0.801687 0.000005 16.1 3.09 0.116 0.511687 0.000004 18.6 2.05
HK17-16 Upper Siwalik 4.22 81.6 46.7 0.783084 0.000008 23.9 4.65 0.118 0.511701 0.000004 18.3 2.06
HK17-18 Upper Siwalik 4.12 72.6 36.3 0.801079 0.000006 26.0 4.86 0.113 0.511552 0.000004 21.2 2.18
HK17-19 Upper Siwalik 4.02 68.4 29.4 0.829414 0.000008 15.6 2.78 0.108 0.511540 0.000003 21.4 2.09
HK17-20 Upper Siwalik 3.93 50.7 47.7 0.764714 0.000008 15.2 2.95 0.117 0.511759 0.000003 17.1 1.96
HK17-21 Upper Siwalik 3.77 61.7 32.4 0.809481 0.000009 11.7 2.14 0.110 0.511674 0.000007 18.8 1.96
HK17-22 Upper Siwalik 3.68 89.3 43.7 0.786630 0.000011 21.4 4.04 0.114 0.511648 0.000003 19.3 2.07
HK17-23 Upper Siwalik 3.56 42.2 20.8 0.790049 0.000010 11.9 2.17 0.110 0.511727 0.000004 17.8 1.88
HK17-24 Upper Siwalik 3.46 68.1 37.1 0.806579 0.000004 12.1 2.21 0.111 0.511613 0.000004 20.0 2.05
HK17-25 Upper Siwalik 3.34 84.3 47.4 0.798762 0.000010 13.3 2.64 0.120 0.511652 0.000004 19.2 2.18

MANDAL ET AL. 128


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

Table 1 (continued)
b b b b b
Age Rb Sr Nd Sm TDM
a 87 86 b 147 144 143 144 b
Samples Formation (Ma) (ppm) (ppm) Sr/ Sr 2 SE (ppm) (ppm) Sm/ Nd Nd/ Nd 2 SE εNd(0) (Ga)

HK17--26 Upper Siwalik 3.26 98.1 38.0 0.818893 0.000007 15.4 3.06 0.120 0.511662 0.000003 19.0 2.16
HK17-27 Upper Siwalik 3.11 83.6 44.6 0.790603 0.000010 12.2 2.49 0.124 0.511727 0.000003 17.8 2.14
HK17-28 Upper Siwalik 3.04 80.5 37.8 0.783998 0.000010 12.4 2.31 0.113 0.511746 0.000003 17.4 1.90
HK17-29 Upper Siwalik 2.85 29.7 13.3 0.767927 0.000014 11.4 2.33 0.123 0.511799 0.000003 16.4 2.03
HK17-30 Upper Siwalik 2.73 56.2 41.2 0.777771 0.000005 13.3 2.62 0.119 0.511763 0.000003 17.1 1.99
HK17-31 Upper Siwalik 2.57 94.2 42.6 0.797562 0.000010 20.2 3.98 0.119 0.511663 0.000003 19.0 2.15
HK17-32 Upper Siwalik 2.49 92.5 35.3 0.834434 0.000013 20.8 4.09 0.119 0.511779 0.000002 16.8 1.97
HK17-33 Upper Siwalik 2.25 39.7 55.2 0.768484 0.000006 14.1 2.79 0.120 0.511728 0.000003 17.8 2.06
HK17-34 Upper Siwalik 2.21 65.1 35.2 0.820125 0.000004 21.7 3.89 0.108 0.511572 0.000002 20.8 2.06
HK17-35 Upper Siwalik 2.11 36.3 13.5 0.764588 0.000009 11.9 2.24 0.114 0.511853 0.000003 15.3 1.77
HK17-36 Upper Siwalik 2.05 34.1 36.8 0.787522 0.000005 15.4 2.88 0.113 0.511530 0.000003 21.6 2.21
HK17-37 Upper Siwalik 1.95 66.3 23.6 0.816887 0.000011 18.2 3.41 0.113 0.511737 0.000003 17.6 1.93
HK17-38 Upper Siwalik 1.79 87.2 34.3 0.827793 0.000009 20.9 3.89 0.112 0.511576 0.000003 20.7 2.13
HK17-39 Upper Siwalik 1.30 29.9 10.7 0.769968 0.000011 12.2 2.19 0.108 0.511823 0.000002 15.9 1.73
HK17-40 Upper Siwalik 0.92 42.4 17.4 0.760282 0.000013 14.9 2.86 0.116 0.511855 0.000003 15.3 1.81
HK17-41 Upper Siwalik 0.67 74.1 36.7 0.785986 0.000012 20.8 3.77 0.110 0.511632 0.000003 19.6 2.01
a b c
Magnetostratigraphic age from Sangode et al. (1996, 1999). Measured in the silicate fractions. The model ages (TDM) were calculated following the model of
DePaolo (1981).

of both sections show some variations, including several excursions, for instance, at ~8.5 Ma in Mohand-Rao
section (Figure 6). The excursion related increases/decreases in εNd(0) values coincide with concomitant
decreases/increases in 87Sr/86Sr ratios. However, a few samples, for example, MR17-04 of depositional age
~8.7 Ma, appear to be decoupled in their isotopic composition. When excluding the short-term excursions,
the εNd(0) values of the Mohand-Rao samples appear to decrease gradually from around 17 at ~9 Ma to
19 at ~6.5 Ma (Figure S7 in the supporting information). Although less clear, the 87Sr/86Sr ratios appear
to increase by about 0.02 over a similar period (Figure S7). Younger sediments from both sections show no
discernible trend in their isotopic compositions.

4.3. Detrital Zircon Geochronology


4.3.1. River Sediments
The Ganga sediments yielded zircon U-Pb ages from ~20 to 2,730 Ma, including three main peaks at ~30 Ma,
~490 Ma, and ~840 Ma (Figure 7). The Yamuna sediments yielded ages from ~38 to 2,630 Ma, with prominent

Figure 4. Scatterplots of major and trace element ratios of sandstones from the Mohand-Rao and Haripur-Khol sections.
(a) Al/Si versus Fe/Si and (b) Zr/Sc versus Th/Sc.

MANDAL ET AL. 129


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

Figure 5. Variation diagram of major and trace elements against Al/Si for samples from the Haripur-Khol and Mohand-Rao sections. (a) Al/Si versus Na2O; (b) Al/Si
versus K2O; (c) Al/Si versus MgO; (d) Al/Si versus Sr; (e) Al/Si versus Rb; (f) Al/Si versus CaO; (g) Al/Si versus Zr; and (h) Al/Si versus TiO2.

age clusters at ~480–980 Ma and a broad peak at 1,880 Ma (Figure 7). In both rivers, Cenozoic (<50 Ma) grains
are characterized by high U/Th ratios (>50; Table S5 in the supporting information). The Ganga sediments are
distinct in showing dominance of Neoproterozoic-Cambrian zircon grains, whereas the Yamuna sediments
are dominated by Paleoproterozoic grains, as shown in previous work (Alizai et al., 2011).
4.3.2. Siwalik Sandstones
Ten samples from the Mohand-Rao section yielded zircon U-Pb ages from ~20 to 3,600 Ma, and 11 samples
from the Haripur-Khol section yielded ages from ~18 to 3,800 Ma. Two main observations can be drawn.
First, regardless of depositional age, the resulting age spectra of both sections are similar and have similar
age clusters, most notably at ~30 Ma, 470–490 Ma, 750–1,050 Ma, 1,800–1,900 Ma, and 2,400–2,600 Ma
(Figures 8 and 9). Second, all samples contain some zircon grains younger than 50 Ma (2%–17%). These grains
have remarkably higher U/Th ratios (12 to 1250) than >50 Ma grains (Tables S6 and S7 in the supporting
information), and they display weak regular zoning to unzoned rims in cathodoluminescence images. Two
samples (MR17-15 and MR17-27) from the Mohand-Rao section and four samples (HK17-16, HK17-32,
HK17-34, HK17-36, and HK17-40) from the Haripur-Khol section contain Early Cretaceous zircon grains with
U-Pb ages of ~105–145 Ma (Figures 8 and 9).

5. Reliability of the εNd(0) Values and 87Sr/86Sr Ratios as Provenance Indicator


For detrital records such as those presented here, changes in mineralogical composition may affect the
whole-rock isotopic composition. For example, low radiogenic Sr isotopic compositions may result from high
proportions of feldspar that have lower Rb/Sr ratios than mica. In other words, changes in 87Sr/86Sr ratios may
reflect changes in mineralogical composition, instead of, or in addition to changes of rock exposure in the
source areas. Most major rock-forming minerals transported from the Himalaya have light rare earth element
(LREE) enriched patterns (e.g., Garzanti et al., 2011). The difference in Nd isotopic composition of these miner-
als is controlled by the age and origin of their source rocks (Goldstein et al., 1984). Accessory minerals such as
zircon, garnet, monazite, allanite, and xenotime have strongly different REE compositions, and therefore, their
variable abundance in clastic sediments may lead to contrasting εNd(0) values (e.g., Garçon et al., 2014). As
group 1 samples generally tend to be richer in detrital zircon (Figure 4b), the more radiogenic Nd isotopic
composition may be controlled by mineralogy. Zircon is typically enriched in heavy rare earth elements
(HREE) (e.g., Garzanti et al., 2011), which leads to high 147Sm/144Nd ratios that, over time, develop high
143
Nd/144Nd ratios. We performed a simple mass balance calculation using the U-Pb age distribution of zircon
grains (section 4.3.2) and the measured bulk silicate Nd and Zr concentration in the analyzed samples, an

MANDAL ET AL. 130


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

87 86
Figure 6. Variations of (a) εNd(0) values, (b) Sr/ Sr and (c) Al/Si ratios in sandstones from the Mohand-Rao and Haripur-Khol sections with time. (bottom) The aver-
87 86
age (±1σ) εNd(0) values and Sr/ Sr ratios of Tethyan Himalayan (THS), Greater Himalayan (GHS), outer-Lesser Himalayan (oLHS), and inner-Lesser Himalayan (iLHS)
sequences shown for comparison (complete data set and their references in supporting information Table S11). Note that the error bars are clipped at the axis
limits. Gray shaded zone = demarcation between iLHS and GHS, THS, and oLHS. (top) Sr and Nd isotopic composition of the modern Ganga and Yamuna river
13
sediments from Singh et al. (2008) and Tripathi et al. (2013). (d) Published δ Csc values of pedogenic carbonate nodules in Siwalik Formation of north-western Indian
Himalaya (Sanyal et al., 2010; Singh et al., 2013; Vögeli et al., 2017). VPDB = Vienna Pee Dee Belemnite.

average zircon 147Sm/144Nd ratio of 0.45, and an average Himalayan zircon Nd concentration of 0.59 ppm
(Garzanti et al., 2010). Results show that a high abundance of zircon grains may increase εNd(0) values by
0.25 to 2.12 ε units. Although such a shift would be more significant for group 1 samples, it cannot
account for the observed offset from the group 2 samples to higher εNd(0) values (Figure S6). The εNd(0)
values of both feldspar- and mica-rich samples broadly overlap with a slight tendency of the feldspar-rich
samples to lower εNd(0) values. The 87Sr/86Sr ratios also overlap without revealing a systematic variation
between feldspar and mica-rich samples (Figure S8). Therefore, the measured Sr-Nd isotopic compositions

MANDAL ET AL. 131


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

can be considered to reflect the source rock rather than variations in the
content of detrital minerals such as mica, feldspar, and zircon.
Covariation of 87Sr/86Sr ratios and εNd(0) values in most samples further
suggests that sediment provenance controls these changes. However,
decoupling of isotopic composition in a few samples, for example,
MR17-04 (Figure 6), may be related to the abundance of other old acces-
sory metamorphic minerals such as monazite and allanite (excursions to
low εNd(0)) or xenotime and garnet (excursions to high εNd(0)), or higher
abundance of old micas with high 87Sr/86Sr ratios.

6. Provenance Constraints From Geochronological and


Isotopic Data
The potential sources of Siwalik sediments exposed along the Mohand-Rao
and Haripur-Khol sections are rocks of the THS, GHS, oLHS, and iLHS. We
compiled published zircon U-Pb ages from these sources in the northwes-
tern Indian and Nepalese Himalayas (Cawood et al., 2007; Colleps et al.,
2018; Ding et al., 2016; Gehrels et al., 2011; Hou et al., 2012; Huang et al.,
2017; Liu et al., 2016, 2017; Mandal et al., 2015, 2016; McKenzie et al., 2011;
Myrow et al., 2003, 2010; Rubatto et al., 2013; Spencer et al., 2012b; Webb
et al., 2011; Zheng et al., 2016; the complete data set is provided in the sup-
porting information Table S10) to create reference age spectra for identifying
the provenance of the studied sediments. This compilation shows that the
THS, GHS, and oLHS rocks have indistinguishable age spectra, except for a
distinctively large ~1,800-Ma peak in the oLHS (Figure 7). Peaks around
500 Ma in the age distribution characterize THS strata and plutons of the
GHS (Gehrels et al., 2011). Approximately 130-Ma zircon grains are reported
from THS strata (Gehrels et al., 2011) as well as from the Late Cretaceous-
Paleocene Singtali Formation rocks of the oLHS (Figure 7; Colleps et al.,
2018). Zircon grains younger than 50 Ma represent Paleogene granitoids
and 17–31 Ma leucogranites and migmatites of the GHS (Figure 7; e.g.,
Aikman et al., 2008; Gao et al., 2016; Liu et al., 2016; Zeng et al., 2011;
Rubatto et al., 2013). Both iLHS- and oLHS-derived zircon grains yield major
peaks in the age distributions around 1,800–1,900 Ma (Figure 7). Potential
sources are structurally cannibalized older foreland basin deposits (e.g., van
der Beek et al., 2006). Detrital zircon age data from the Late Paleocene to
mid-Miocene foreland basin deposits in northwestern India have age distri-
butions representative of THS, GHS, and oLHS sources (Figure 7; Ravikant
et al., 2011). In summary, the overlapping age distribution in Himalayan tec-
tonostratigraphic units confound clear identification of provenance of the
foreland basin deposits based on zircon age spectra alone (cf. Ravikant
et al., 2011), which should be considered in future provenance studies.
Our compilation of the published bedrock Sr-Nd isotopic data (Ahmad
et al., 2000; Imayama & Arita, 2008; Mandal et al., 2015; McKenzie et al.,
2011; Martin et al., 2005; Miller et al., 2001; Parrish & Hodges, 1996;
Richards et al., 2005; Robinson et al., 2001; the complete data set is pro-
Figure 7. Age histograms and normalized probability density plots (PDP) of
zircon U-Pb age populations for modern Ganga and Yamuna river sediments vided in the supporting information Table S11) suggests that THS rocks
(this study). (bottom) Reference populations from various tectonostrati- have an average εNd(0) value and 87Sr/86Sr(0) ratio of 15.1 (1σ = 3.5)
graphic units of Himalaya (see supporting information for the compiled and 0.761 (1σ = 0.017), respectively. GHS rocks yield an average εNd(0)
data set and their references) and (middle) age populations from Late value of 15.2 (1σ = 2.7) and 87Sr/86Sr(0) ratio of 0.766 (1σ = 0.037),
Paleocene to mid-Miocene Sub-Himalayan foreland basin deposits in NW
whereas the oLHS rocks have a mean εNd(0) value of 14.6 (1σ = 3) and
India (Colleps et al., 2018; Ravikant et al., 2011). n = number of individual
zircon U-Pb ages plotted as PDP. THS = Tethyan Himalayan Sequence;
87
Sr/86Sr(0) ratio of 0.787 (1σ = 0.046). In contrast, more negative εNd(0)
GHS = Greater Himalayan Sequence; LHS = Lesser Himalayan Sequence. values and high 87Sr/86Sr(0) ratios, with mean values of 23.4 (1σ = 2.4)
and 0.905 (1σ = 0.136), respectively, are characteristic of iLHS rocks

MANDAL ET AL. 132


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

(Figure 6). This isotopic fingerprint of the iLHS rocks in conjunction with
~1,800–1,900 Ma zircon age distribution peaks will therefore help recog-
nizing their denudation in the hinterland and input into the foreland basin.

7. Provenance Interpretation
7.1. Detrital Zircon Record
The detrital zircon U-Pb age distributions in the Siwalik sediments from the
Mohand-Rao and Haripur-Khol sections, compared to the reference age
spectra of various source units (Figure 7), suggest that the sediments
originated from a combination of any of the THS, GHS, and low-grade
equivalents of GHS (e.g., Haimanta Formation; Figure 2a), oLHS, and iLHS
sources. In the Siwaliks of both sections, the abundance of Oligocene-
Miocene zircon grains suggests sediment derivation from the GHS leuco-
granites and migmatites (e.g., Searle et al., 1999; Weinberg, 2016, and
references therein). Furthermore, the occurrence of Early Cretaceous
zircon grains suggests that Late Cretaceous-Paleocene Singtali
Formation rocks of the oLHS may have been a source of the studied
Siwalik sediment (Colleps et al., 2018; Webb et al., 2011). The dominance
of Neoproterozoic-Cambrian zircon grains over Paleoproterozoic grains
suggests a smaller contribution from the iLHS sources, throughout the
sedimentation record. Unfortunately, the detrital zircon age patterns can-
not serve as evidence for or against the occurrence of recycled sediments
from the older Sub-Himalayan foreland basin deposits. However, exclusive
sediment recycling from the older foreland basin deposits into younger
Siwaliks of the Mohand-Rao and Haripur-Khol sections can be ruled out.
Because of the occurrence of Oligocene-Miocene detrital zircon grains in
the Siwaliks of both sections, which do not occur in the Late Paleocene
to mid-Miocene Sub-Himalayan foreland basin deposits (Figure 7; Baral
et al., 2016; DeCelles et al., 1998; Ravikant et al., 2011). Furthermore,
cannibalization of older Sub-Himalayan foreland basin deposits would
deliver dominantly Neoproterozoic-Cambrian zircon grains (Figure 7).
Therefore, the recycling of old Sub-Himalayan foreland basin material
would yield an overall upsection increase in Neoproterozoic-Cambrian
grains, which is not observed in our data set.
The detrital zircon U-Pb ages obtained from the present-day Ganga
sediment point to GHS and THS being the dominant sources, whereas
contributions from the Lesser Himalayan sequences (i.e., iLHS and
oLHS) seem to be subsidiary. In contrast, the results of the present-
day Yamuna sediment suggest higher contributions from the Lesser
Himalayan (i.e., iLHS and oLHS) sources than from combinations of the
THS and GHS sources (Figure 7). In both populations, the diagnostic
Oligocene-Miocene zircon grains indicate sediment derivation from
GHS migmatites and leucogranites, which are exposed in their head-
waters (Figure 7). The large share of Lesser Himalayan Sequence-derived
zircon grains in the Yamuna population fits with the larger Lesser
Himalayan exposure area in the Yamuna catchment (66%) compared
to the Ganga catchment (55%). Therefore, we argue that the zircon
age populations of the present-day Ganga and Yamuna sediments
Figure 8. Age histograms and normalized probability density plots of zircon
U-Pb age populations for sandstones from the Mohand-Rao section.
reflect source rock distributions within the catchment.
Numbers in the top left- depositional age. See Figure 3 for the stratigraphic In summary, the detrital zircon U-Pb age spectra suggest that the
position of the samples. Siwalik sediments of the Mohand-Rao and Haripur-Khol sections were
derived from THS, GHS, oLHS, and iLHS sources. Owing to the overlap-
ping zircon U-Pb ages of the THS, GHS, and oLHS, it is, however,

MANDAL ET AL. 133


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

difficult to further separate their relative contributions to the Late


Cenozoic foreland basin-fill sediments. Variable mixtures of THS/GHS/
oLHS source contributions consistently dominated the detrital popula-
tions of both sections.

7.2. Bulk-Sediment Geochemical and Isotopic Record


The group 2 sediments of the Mohand-Rao and Haripur-Khol sections
are characterized by relatively uniform geochemical composition
(Figure 4), which implies that their source remained rather unchanging
over the past ~10 Myr. This does not necessarily imply sediment input
from any particular source, but merely a relatively constant contribu-
tion from several sources (e.g., Galy et al., 2010). Because albite
behaves like potassic feldspar during sediment transport and deposi-
tion (Garzanti et al., 2010), its higher influence in eight samples
(MR17-19, MR17-28, MR17-31, HK17-17, HK17-24, HK17-30, HK17-33,
and HK17-36) may indicate derivation from a different source. The
low (<19) εNd(0) values of some of these samples hint to LHCS
sources. The pronounced shift in isotopic composition toward negative
εNd(0) values reflect an old crust contribution, that is, detritus eroded
from the iLHS. Conversely, a shift towards higher εNd(0) values reflects
an increase in detritus eroded either from more juvenile source
regions such as the THS, GHS, and oLHS (Figure 6; Ahmad et al.,
2000; Parrish & Hodges, 1996; Robinson et al., 2001; Richards et al.,
2005) or recycled from older Sub-Himalayan foreland basin deposits
(Figure 10; Najman et al., 2000). Note that old and young refer to
the average protolith age, not to the timing of metamorphic overprint.
The anomalies to high εNd(0) values are mainly associated with group 1
sediments of the Haripur-Khol section (Figure 6). Interestingly, the Sr-Nd
isotopic composition of these sediments overlaps with the composition
of Late Oligocene to mid-Miocene Dagshai and Kasauli sediments
(Figure 10), which suggests that group 1 sediments contain a significant
amount of recycled material from older Sub-Himalayan foreland basin
deposits. This erosional recycling of older foreland basin-fill deposits
into younger Siwalik sediments may explain the noisy Sr-Nd isotopic sig-
nals in the time series of our data set. As the compositional anomalies
are mainly spikes in our data set, the excursions to lower εNd(0) values
may reflect landslides in the source areas that are either swamping the
river with iLHS-derived material or that are blocking sediments from
other sources. Alternatively, this effect could be due to the varying influ-
ences of tributaries that preferentially drain iLHS units.
Excluding the discussed anomalies, the direction of the Nd and Sr isotopic
shift between ~9 and 6.5 Ma (Figure S7) suggests increased flux from the
iLHS, implying a gradually increasing erosional unroofing of the inner-
Lesser Himalayan units. The isotopic compositions of the Siwalik sedi-
ments from the Mohand-Rao and Haripur-Khol sections fall between the
compositional space of the GHS/THS/oLHS and iLHS, with most samples
being closer to the GHS/THS/oLHS field (Figure 10). Source apportionment
modeling using a Monte Carlo approach (details in supporting informa-
tion; Awasthi et al., 2018) and assuming two-component mixing between
the GHS (as GHS/THS/oLHS rocks have overlapping ranges in isotopic
Figure 9. Age histograms and normalized probability plots of zircon U-Pb composition) and iLHS indicate that GHS/THS/oLHS rocks account for
age populations for sandstones from the Haripur-Khol section. Same >60% of the detritus deposited in the foreland basin over the past
legend as in Figure 8. ~10 Myr (Figure 10).

MANDAL ET AL. 134


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

Figure 10. Sr-Nd isotope mixing plot showing the composition of the Mohand-Rao and Haripur-Khol sediments deposited since ~10 Ma along with compositional
fields of major tectonostratigraphic units of northwestern Indian Himalaya. Rectangles are the distribution of isotopic composition in the Himalayan tectonostrati-
graphic units represented as average (±1σ) values calculated from the compiled published data set (complete data set and references in the supporting
information Table S11). Sr-Nd isotopic data of the Sub-Himalayan Late Oligocene to mid-Miocene Dagshai and Kasauli sediments compiled from Najman et al. (2000).
The dots in the background, color coded on the basis of mixing proportions between two end-members (i.e., GHS and iLHS), show the results of Monte Carlo
simulations (n = 100,000) that predict the Sr-Nd isotopic budget of bulk sediments estimated on the basis of binary end-member sources. Light orange region is the
predicted bulk-sediment isotopic composition using 90% GHS-derived and 10% iLHS-derived materials. Light purple region is the predicted bulk-sediment
isotopic composition using 70% GHS-derived and 30% iLHS-derived materials. GHS = Greater Himalayan Sequence; THS = Tethyan Himalayan Sequence;
iLHS = inner-Lesser Himalayan Sequence; oLHS = outer-Lesser Himalayan Sequence.

7.3. Summary of Provenance Interpretation


Our provenance analysis using both detrital zircon U-Pb geochronology and bulk silicate Sr-Nd isotope geo-
chemistry suggests that the source regions for the Siwalik sediments of the Mohand-Rao and Haripur-Khol
sections were compositionally relatively constant. The data further suggest that during the past ~10 Myr,
the foreland basin sediments have been dominated by material derived from the GHS/THS/oLHS sources.
We interpret the long-term changes in the Sr-Nd isotopic compositions of the Mohand-Rao sediments
between ~9 and 6.5 Ma to indicate increasing erosional unroofing of iLHS. However, the subsequent sedi-
ments show no systematic evolution of the isotopic composition that could be attributed to increasing pro-
portions of iLHS-derived material with time. Instead, the combined geochemical and isotopic fingerprints of
group 1 samples reveal significant erosional recycling of older Sub-Himalayan foreland basin deposits since
the Late Miocene (~5.5 Ma).

8. Discussion
8.1. Source-To-Sink Connection During Siwalik Deposition
Our results support the existence of a relatively steady drainage system in the Dehradun reentrant area dur-
ing the Late Cenozoic. This differs from the Siwalik record of the Kangra reentrant area, located ~200 km
northwest of our study area (Figure 1), where an abrupt disappearance of GHS-derived material at ~7 Ma
is attributed to a tectonically driven drainage diversion in the hinterland (Najman et al., 2009). The continuous
flux of GHS/THS-derived material to the Siwaliks of the Mohand-Rao and Haripur-Khol sections supports
earlier studies inferring spatial and temporal stability of Himalayan drainages since the inception of the
Main Boundary Thrust, possibly during the Late Miocene (Gupta, 1997). While progressive exhumation of
structurally deeper units and different contributions of tributaries can be invoked to explain the temporal
changes in the various zircon age peak heights (Alizai et al., 2011; DeCelles et al., 1998), such processes

MANDAL ET AL. 135


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

apparently do not affect the zircon age ranges in our time series of
Himalayan-derived sediments. This assertion is supported by the similar
zircon geochronology results that we obtained from both the Siwaliks
and modern river-borne sediments. The similar zircon age ranges in mod-
ern Ganga and Yamuna river sediments pose a major difficulty in distin-
guishing their paleosediments in the Siwaliks of both studied sections.
Given the geographic location of the Mohand-Rao site, nearly half way
between the outlets where the Yamuna and Ganga rivers debouch into
the foreland basin (Figure 2), the sediments may have been deposited
by either one in the past. The thick multistoried sandstone complex of
the Mohand-Rao section, with its dominant sheet geometries, planar to
trough cross-stratified beds, low mudstone contents, the occurrence of
frequent erosional events, and consistent paleoflow indicates sediment
deposition by a braided river in the distal part of an alluvial fan (Kumar
et al., 2004). Stratigraphic analyses along the western half of the Mohand
anticline identified two major drainage systems of coalescing alluvial
megafan type, established at the proximal basin border by ~10 Ma
(Kumar, 1993; Kumar et al., 2004). Between ~9 and 6 Ma, the prevalence
of iLHS-derived material and the simultaneous occurrence of >15-m-thick
channel sandstones with dominantly southward paleoflow in the Siwaliks
of the Mohand-Rao section suggest sediment accumulation by a large
transverse Himalayan river. We hypothesize that during that period both
the Yamuna and Ganga paleo rivers may have contributed to the
Mohand-Rao site, probably in a fan-marginal position, although not neces-
sarily at the same time (Figure 11).
Starting at ~6.5 Ma, the εNd(0) values in the Mohand-Rao sediments
suggest reduced sediment influx from the iLHS. In the Siwaliks of
the Kangra reentrant area (Figure 2), Najman et al. (2009) observed
an increasing influx of mafic volcanic and dolomite lithic fragments
along with Proterozoic white micas of iLHS origin at a similar time
(~6 Ma), which they ascribed to tectonically induced exhumation of
the metamorphosed iLHS (i.e., the LHCS). Interestingly, the period of
diminishing iLHS contribution in the Mohand-Rao sediments was fol-
lowed by an abrupt sandstone-conglomerate facies transition at
~4.8 Ma (Figure 3). Within Himalayan rivers that feed the Gangetic
plain, abrupt grain size coarsening is frequently observed to occur at
a distance of, ~8- to 20-km downstream from the MFT, as a result of
sorting processes during selective deposition of the sediment load
(Dingle et al., 2017; Dubille & Lavé, 2015). We thus interpret the
appearance of thick conglomerates to indicate the southward migra-
tion of the gravel front in response to thrust wedge advance and
shortening of the foreland basin (Dubille & Lavé, 2015). This inference
is consistent with previous studies of conglomerate clasts in the
Mohand-Rao section (Kumar et al., 2003). The arrival of the thrust front
near the depositional site would have increased the likelihood that
more proximal sources (i.e., oLHS rocks) directly supply sediments to
Figure 11. Possible scenarios for the Late Miocene-Pleistocene evolution of the Mohand-Rao site via small rivers. We thus suggest that since
the Ganga and Yamuna drainage system in the northwestern Indian ~6.5 Ma, the source-to-sink connection at the Mohand-Rao site was
Himalayan foreland basin, constructed using provenance studies of the accomplished by small piedmont rivers along with probably lesser con-
Mohand-Rao and Haripur-Khol sections. Note the different positions of the
tributions from the Yamuna and Ganga paleo rivers (Figure 11).
depositional sites with respect to the advancing thrust front.
The persistent occurrence of GHS/THS-derived material in the Haripur-
Khol section likewise testifies to sediment accumulation by a transverse
Himalayan river. The diachronous sandstone-conglomerate transition in

MANDAL ET AL. 136


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

Figure 12. Schematic reconstruction of the Late Miocene to Pleistocene Himalayan fold-and-thrust belt in northwestern
India. Same colors of the tectonostratigraphic units as in Figure 2. Same abbreviations as in Figures 2 and 6.

the two studied sections (i.e., ~4.8 Ma in Mohand-Rao versus <0.7 Ma in Haripur-Khol) over a horizontal
distance of ~60 km indicates a strong obliquity between timelines and facies limits, which in turn suggests
that the Plio-Pleistocene Siwalik sediments of the Haripur-Khol section were deposited in a more distal
alluvial fan setting than the Mohand-Rao section (Dubille & Lavé, 2015). The Giri River is the southernmost
hinterland draining tributary of the Yamuna River, which it presently joins in the Dehradun reentrant
(Figure 2). However, during the Plio-Pleistocene, before initiation of the MFT, these two rivers may have
been feeding the foreland basin separately. Therefore, these rivers could have formed their own fans, and
the Haripur-Khol section may thus record the deposits of the Giri River. However, the provenance of the
Haripur-Khol sediments, especially the occurrence of iLHS-derived material, is inconsistent with sediments
deposited only by the Giri River, because the iLHS is not exposed in its catchment (Figure 2). The
dominantly southward directed paleoflow further refutes the possibility of sediment accumulation by
other transverse rivers such as the Sutlej to the west. Therefore, we suggest that the Siwaliks of the
Haripur-Khol section represent deposits of the Yamuna River.
8.2. Implications for the Denudation and Tectonic History of the Hinterland
Provenance data presented here support previous studies inferring that the Paleoproterozoic iLHS rocks
became exposed during the Late Miocene (Colleps et al., 2018; DeCelles et al., 1998; Huyghe et al., 2001,
2005; Najman et al., 2009, 2010). Our results provide a clear indication of iLHS erosion at ~9 Ma. The ~1–
2 Ma asynchronicity in the εNd(0) signatures of iLHS erosion across the Siwaliks of northwestern Indian
and Nepalese Himalayas does not imply diachronous exposure and erosion of this unit. Varying tributary con-
tributions and swamped εNd(0) signal in the bulk rock analyses by dilution with the more juvenile source (i.e.,
THS/GHS/oLHS) may explain the diachronicity (e.g., Najman et al., 2009). Therefore, ~9 Ma should not be
stringently considered as the period of initial iLHS unroofing across the Kumaun-Garhwal Himalaya.

MANDAL ET AL. 137


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

The appearance of LHCS-derived material in the Siwalik sediments of the Mohand-Rao and Haripur-Khol sec-
tions at ~6.6 Ma (section 7.2) is consistent with previous studies linking the initial unroofing of this sequence
to out-of-sequence movement along the Ramgarh-Munsiari Thrust (Figure 12; Célérier, Harrison, Beyssac,
et al., 2009; Stübner et al., 2018; Vannay et al., 2004). Najman et al. (2009) also noted a contemporaneous
appearance of LHCS-derived material in the Siwaliks of the Kangra reentrant (Figure 1). Accordingly, the wide-
spread Late Miocene-Pleistocene erosion of LHCS most likely heralds the earliest record of a basement thrust
ramp (Caldwell et al., 2013).
Our results are compatible with previous detrital apatite fission track (AFT) thermochronology data from
the Siwalik sediments of the Nepalese Himalaya, which indicate erosional recycling of older Sub-
Himalayan foreland basin deposits into the Plio-Pleistocene strata. This may reflect the enlargement of
the Himalayan orogenic wedge through frontal accretion of older foreland basin strata and concomitant
cannibalization of proximal foreland basin depositional units. This interpretation was based on the ~5-Ma
difference between the thermochronological and depositional ages of a 1 ± 0.5-Ma sample (van der Beek
et al., 2006). Our findings provide evidence for the onset of significant erosional recycling of older fore-
land basin-fill material into younger Siwaliks at ~5.5 Ma, which may indicate when thrusting had propa-
gated from the Lesser Himalaya into the foreland basin. This age for the onset of Sub-Himalayan thrusting
agrees with previous AFT thermochronological data from the Early to mid-Miocene foreland basin-fill
strata of the Subathu salient (Figure 1), which yielded AFT ages of 5.3 ± 0.6 Ma (Najman et al., 2004).
Our findings of an overall weak iLHS source contribution to the Siwalik sediments of the Mohand-Rao and
Haripur-Khol sections concur with similar studies on both the Bay of Bengal sediments and the Siwaliks of
the central Himalaya in Nepal (Galy et al., 2010; Huyghe et al., 2001, 2005). Furthermore, the provenance of
these Late Miocene to Pleistocene sediments does not differ perceptibly from the bulk of the Himalayan
material that is nowadays shed to the foreland basin via the Yamuna and Ganga rivers (Figure 6). Our results,
therefore, support earlier studies inferring that Himalayan erosion has remained relatively unchanged since
at least the Late Miocene (Bernet et al., 2006; Galy et al., 2010; van der Beek et al., 2006). The estimated
erosion/exhumation rates are roughly an order of magnitude higher in the topographically high, northern
part of the Himalaya, where the GHS and THS units are dominantly exposed, compared to the low, southern
part exposing the iLHS and oLHS units (Morell et al., 2017; Scherler et al., 2014). This topographic pattern and
the corresponding erosion/exhumation rate gradients are attributed to a thrust ramp at a midcrustal level
(Caldwell et al., 2013). The observation that GHS and THS source contributions have overwhelmed the Late
Cenozoic Himalayan-derived sediments suggests, therefore, that this inferred ramp exists since at least the
Late Miocene. The actual position of the active ramp, however, may have changed through time as new
thrust-bounded slivers of crust were accreted at the base of the southward progressing orogenic wedge
(e.g., Mercier et al., 2017).

8.3. Late Cenozoic Climatic Forcing of Himalayan-Derived Sediments


Studies of paleoclimatic reconstructions of the Himalayan foreland basin have documented a dramatic
shift in the stable carbon isotopic composition of pedogenic carbonates towards heavier components
(δ13C) around ~7 Ma (Figure 6d; e.g., Quade & Cerling, 1995; Sanyal et al., 2010; Vögeli et al., 2017).
This Late-Miocene event has been interpreted to reflect the vegetation shifting from dominantly C3 plant
types (i.e., mostly trees) to more C4-dominated (i.e., mainly grassland). Since the pCO2 did not change
drastically during the Miocene (e.g., Pagani et al., 2005), the expansion of C4 species has been attributed
to climate change, that is, less abundant and/or more seasonally concentrated precipitations. The Sr-Nd
isotopic composition of the Mohand-Rao sediments shows no systematic shift during the Late Miocene
(Figure 6), implying that this climatic change did not significantly influence the spatial distribution of ero-
sion, and hence sediment provenance in our study area. This observation agrees with earlier studies on
the Bay of Bengal sediments (Galy et al., 2010). The tectonic forcing appears to exert the primary control
on Himalayan erosion (e.g., Godard et al., 2014; Scherler et al., 2014).

9. Conclusions
We provided new Sr-Nd isotopic compositions, detrital zircon U-Pb age spectra, and major and trace ele-
ment compositions along two sections of the Siwalik sediments dated ~9.8 to <0.7 Ma. Comparison of

MANDAL ET AL. 138


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

the detrital zircon U-Pb age spectra and the bulk silicate Sr-Nd isotopic compositions with those of the
main Himalayan tectonostratigraphic source units suggests stable erosion of the Himalayan orogenic belts
since the Late Miocene. Our data indicate that erosion of the LHCS started around 6 Ma, possibly with
out-of-sequence movement along the Ramgarh-Munsiari Thrust. Reworking of older foreland basin depos-
its since ~5.5 Ma suggests that thrusting had propagated from the Lesser Himalaya into the foreland
basin during the Late Miocene. The detritus in the studied northwestern part of the Himalayan foreland
basin was supplied by southward directed transverse rivers, analogous to rivers feeding the foreland basin
at present. We did not identify any causal link between the Late Miocene climate change and the result-
ing mass flux from the orogen. Although the iLHS has been exposed and eroded since at least the Late
Miocene, its contribution to the Siwaliks of the Mohand-Rao and Haripur-Khol sections remains small
compared to the much larger contributions from the GHS and THS sources. This observation suggests
that the high topography and rock uplift rate in the Greater and Tethyan Himalaya persisted during
the Siwalik deposition. This supports the notion that duplexing of the Lesser Himalayan lithotectonic zone
is active since at least ~10 Ma.

Acknowledgments References
This work has been supported by an
Alexander von-Humboldt postdoctoral Ahmad, T., Harris, N., Bickle, M., Chapman, H., Bunbury, J., & Prince, C. (2000). Isotopic constraints on the structural relationships between the
fellowship awarded to S. K. Mandal. D. lesser Himalayan series and the high Himalayan crystalline series, Garhwal Himalaya. Geological Society of America Bulletin, 112(3),
Scherler was supported by the DFG 467–477. https://doi.org/10.1130/0016-7606(2000)112<467:ICOTSR>2.0.CO;2
(Deutsche Forschungsgemeinschaft) Aikman, A. B., Harrison, T. M., & Lin, D. (2008). Evidence for early (> 44 Ma) Himalayan crustal thickening, Tethyan Himalaya, southeastern
grant SCHE1676/3-1. Field support from Tibet. Earth and Planetary Science Letters, 274(1-2), 14–23. https://doi.org/10.1016/j.epsl.2008.06.038
P. K. Mukherjee and R. Kumar (former Alizai, A., Carter, A., Clift, P. D., VanLaningham, S., Williams, J. C., & Kumar, R. (2011). Sediment provenance, reworking and transport processes
scientist, Wadia Institute of Himalayan in the Indus River by U–Pb dating of detrital zircon grains. Global and Planetary Change, 76(1-2), 33–55. https://doi.org/10.1016/j.
Geology, Dehradun) is gratefully gloplacha.2010.11.008
acknowledged. Special thanks to R. Allen, P. A. (2017). Sediment routing systems: The fate of sediment from source to sink. Cambridge: Cambridge Univ. Press. https://doi.org/
Kapannusch for his assistance in the 10.1017/9781316135754
field. We are grateful to A. Mohammadi Allen, P. A., & Allen, J. R. (2013). Basin analysis: Principles and application to petroleum play assessment. Chichester, UK: John Wiley & Sons.
for his help on the zircon separation, O. Avouac, J.-P. (2003). Mountain building, erosion, and the seismic cycle in the Nepal Himalaya. Advances in Geophysics, 46, 1–80. https://doi.
Laurent and J. Sliwinski for their support org/10.1016/S0065-2687(03)46001-9
with the LA-ICP-MS, K. Kunze and S. Awasthi, N., Ray, E., & Paul, D. (2018). Sr and Nd isotope compositions of alluvial sediments from the Ganga Basin and their use as potential
Mayanna for their assistance with proxies for source identification and apportionment. Chemical Geology, 476, 327–339. https://doi.org/10.1016/j.chemgeo.2017.11.029
cathodoluminescence and backscatter Baral, U., Lin, D., & Chamlagain, D. (2016). Detrital zircon U–Pb geochronology of the Siwalik Group of the Nepal Himalaya: Implications for
imaging, A. Gottsche, K. Hahne, and H. provenance analysis. International Journal of Earth Sciences, 105(3), 921–939. https://doi.org/10.1007/s00531-015-1198-7
Rothe for their assistance with X-ray Bernet, M., van der Beek, P., Pik, R., Huyghe, P., Mugnier, J. L., Labrin, E., & Szulc, A. (2006). Miocene to recent exhumation of the central
fluorescence and trace element Himalaya determined from combined detrital zircon fission-track and U/Pb analysis of Siwalik sediments, western Nepal. Basin Research,
analysis, B. Hübner for her assistance 18(4), 393–412. https://doi.org/10.1111/j.1365-2117.2006.00303.x
206 238
with Sr-Nd isotopic analysis, and J. Black, L. P., Kamo, S. L., Allen, C. M., Davis, D. W., Aleinikoff, J. N., Valley, J. W., et al. (2004). Improved Pb/ U microprobe geochronology by
Glodny for his assistance with mineral the monitoring of a trace-element-related matrix effect; SHRIMP, ID–TIMS, ELA–ICP–MS and oxygen isotope documentation for a series of
separation. Furthermore, we thank J. zircon standards. Chemical Geology, 205, 115–140. https://doi.org/10.1016/j.chemgeo.2004.01.003
Schüssler for insightful discussions and Bollinger, L., Avouac, J. P., Beyssac, O., Catlos, E. J., Harrison, T. M., Grove, M., et al. (2004). Thermal structure and exhumation history of the
suggestions. Thorough and constructive Lesser Himalaya in central Nepal. Tectonics, 23, TC5015. https://doi.org/10.1029/2003TC001564
comments by Y. Najman and two Brozovic, N., & Burbank, D. W. (2000). Dynamic fluvial systems and gravel progradation in the Himalayan foreland. GSA Bulletin, 112(3),
anonymous reviewers, as well as the 394–412. https://doi.org/10.1130/0016-7606(2000)112<394:dfsagp>2.0.co;2
Associate Editor P. van der Beek helped Burchfiel, B. C., Zhiliang, C., Hodges, K. V., Yuping, L., Royden, L. H., Changrong, D., & Jiene, X. (1992). The South Tibetan detachment system,
us to improve the manuscript. T. Himalayan Orogen: Extension contemporaneous with and parallel to shortening in a collisional mountain belt. In B. C. Burchfiel, C.
Schildgen is acknowledged for the Zhiliang, K. V. Hodges, L. Yuping, L. H. Royden, D. Changrong, & Xujiene (Eds.), The South Tibetan detachment system, Himalayan Orogen:
efficient editorial handling of this Extension contemporaneous with and parallel to shortening in a collisional mountain belt, Geological Society of America Special Papers, (Vol.
manuscript. All the presented data are 269, pp. 1–41). https://doi.org/10.1130/SPE269-p1
available in the supporting information. Burg, J. P., & Chen, G. M. (1984). Tectonics and structural zonation of southern Tibet, China. Nature, 311(5983),
219. https://doi.org/10.1038/311219a0–223
Caldwell, W. B., Klemperer, S. L., Lawrence, J. F., & Rai, S. S. (2013). Characterizing the Main Himalayan Thrust in the Garhwal Himalaya, India
with receiver function CCP stacking. Earth and Planetary Science Letters, 367, 15–27. https://doi.org/10.1016/j.epsl.2013.02.009
Catlos, E., Harrison, T. M., Kohn, M. J., Grove, M., Ryerson, F., Manning, C. E., & Upreti, B. (2001). Geochronologic and thermobarometric
constraints on the evolution of the Main Central Thrust, central Nepal Himalaya. Journal of Geophysical Research, 106(B8), 16,177–16,204.
https://doi.org/10.1029/2000JB900375
Cawood, P. A., Johnson, M. R. W., & Nemchin, A. A. (2007). Early Palaeozoic orogenesis along the Indian margin of Gondwana: Tectonic
response to Gondwana assembly. Earth and Planetary Science Letters, 255(1–2), 70–84. https://doi.org/10.1016/j.epsl.2006.12.006
Célérier, J., Harrison, T. M., Beyssac, O., Herman, F., Dunlap, W. J., & Webb, A. A. G. (2009). The Kumaun and Garwhal Lesser Himalaya, India:
Part 2. Thermal and deformation histories. GSA Bulletin, 121(9-10), 1281–1297. https://doi.org/10.1130/b26343.1
Célérier, J., Harrison, T. M., Webb, A. A. G., & Yin, A. (2009). The Kumaun and Garwhal Lesser Himalaya, India: Part 1. Structure and stratigraphy.
Geological Society of America Bulletin, 121(9-10), 1262–1280. https://doi.org/10.1130/b26344.1
Clift, P. D. (2017). Cenozoic sedimentary records of climate-tectonic coupling in the Western Himalaya. Progress in Earth and Planetary
Science, 4(1). https://doi.org/10.1186/s40645-017-0151-8
Colleps, C. L., McKenzie, N. R., Stockli, D. F., Hughes, N. C., Singh, B. P., Webb, A. A. G., et al. (2018). Zircon (U-Th)/He thermochronometric
constraints on Himalayan Thrust Belt exhumation, bedrock weathering, and cenozoic seawater chemistry. Geochemistry, Geophysics,
Geosystems, 19, 257–271. https://doi.org/10.1002/2017GC007191

MANDAL ET AL. 139


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

DeCelles, P., Carrapa, B., Gehrels, G., Chakraborty, T., & Ghosh, P. (2016). Along-strike continuity of structure, stratigraphy, and kinematic
history in the Himalayan thrust belt: The view from northeastern India. Tectonics, 35, 2995–3027. https://doi.org/10.1002/2016TC004298
DeCelles, P., Gehrels, G., Quade, J., Ojha, T., Kapp, P., & Upreti, B. (1998). Neogene foreland basin deposits, erosional unroofing, and the
kinematic history of the Himalayan fold-thrust belt, western Nepal. Geological Society of America Bulletin, 110(1), 2–21. https://doi.org/
10.1130/0016-7606(1998)110<0002:nfbdeu>2.3.co;2
DeCelles, P. G., Robinson, D. M., Quade, J., Ojha, T., Garzione, C. N., Copeland, P., & Upreti, B. N. (2001). Stratigraphy, structure, and tectonic
evolution of the Himalayan fold-thrust belt in western Nepal. Tectonics, 20(4), 487–509. https://doi.org/10.1029/2000TC001226
DePaolo, D. J. (1981). Neodymium isotopes in the Colorado Front Range and crust–mantle evolution in the Proterozoic. Nature, 291(5812),
193. https://doi.org/10.1038/291193a0–196
87 86
Derry, L. A., & France-Lanord, C. (1996). Neogene Himalayan weathering history and river Sr Sr: Impact on the marine Sr record. Earth and
Planetary Science Letters, 142(1-2), 59–74. https://doi.org/10.1016/0012-821x(96)00091-x
Ding, H., Zhang, Z., Dong, X., Tian, Z., Xiang, H., Mu, H., et al. (2016). Early Eocene (c. 50 Ma) collision of the Indian and Asian continents:
Constraints from the North Himalayan metamorphic rocks, southeastern Tibet. Earth and Planetary Science Letters, 435, 64–73. https://doi.
org/10.1016/j.epsl.2015.12.006
Dingle, E. H., Attal, M., & Sinclair, H. D. (2017). Abrasion-set limits on Himalayan gravel flux. Nature, 544(7651),
471. https://doi.org/10.1038/nature22039–474
Dubille, M., & Lavé, J. (2015). Rapid grain size coarsening at sandstone/conglomerate transition: Similar expression in Himalayan modern
rivers and Pliocene molasse deposits. Basin Research, 27(1), 26–42. https://doi.org/10.1111/bre.12071
Galy, V., France-Lanord, C., Peucker-Ehrenbrink, B., & Huyghe, P. (2010). Sr–Nd–Os evidence for a stable erosion regime in the Himalaya
during the past 12 Myr. Earth and Planetary Science Letters, 290(3-4), 474–480. https://doi.org/10.1016/j.epsl.2010.01.004
Gansser, A. (1964). Geology of the Himalayas. London: Interscience Publishers.
Gao, L.-E., Zeng, L., Gao, J., Shang, Z., Hou, K., & Wang, Q. (2016). Oligocene crustal anatexis in the Tethyan Himalaya, southern Tibet. Lithos,
264, 201–209. https://doi.org/10.1016/j.lithos.2016.08.038
Garçon, M., Chauvel, C., France-Lanord, C., Limonta, M., & Garzanti, E. (2014). Which minerals control the Nd–Hf–Sr–Pb isotopic compositions
of river sediments? Chemical Geology, 364, 42–55. https://doi.org/10.1016/j.chemgeo.2013.11.018
Garzanti, E., Andó, S., France-Lanord, C., Censi, P., Vignola, P., Galy, V., & Lupker, M. (2011). Mineralogical and chemical variability of fluvial
sediments 2. Suspended-load silt (Ganga–Brahmaputra, Bangladesh). Earth and Planetary Science Letters, 302(1-2), 107–120. https://doi.
org/10.1016/j.epsl.2010.11.043
Garzanti, E., Andò, S., France-Lanord, C., Vezzoli, G., Censi, P., Galy, V., & Najman, Y. (2010). Mineralogical and chemical variability of fluvial
sediments: 1. Bedload sand (Ganga–Brahmaputra, Bangladesh). Earth and Planetary Science Letters, 299(3-4), 368–381. https://doi.org/
10.1016/j.epsl.2010.09.017
Gehrels, G., Kapp, P., DeCelles, P., Pullen, A., Blakey, R., Weislogel, A., et al. (2011). Detrital zircon geochronology of pre-Tertiary strata in the
Tibetan-Himalayan orogen. Tectonics, 30, TC5016. https://doi.org/10.1029/2011TC002868
Godard, V., Bourlès, D. L., Spinabella, F., Burbank, D. W., Bookhagen, B., Fisher, G. B., et al. (2014). Dominance of tectonics over climate in
Himalayan denudation. Geology, 42(3), 243–246. https://doi.org/10.1130/g35342.1
Goldstein, S., O’nions, R., & Hamilton, P. (1984). A Sm-Nd isotopic study of atmospheric dusts and particulates from major river systems. Earth
and Planetary Science Letters, 70(2), 221–236. https://doi.org/10.1016/0012-821x(84)90007-4
Guillong, M., von Quadt, A., Sakata, S., Peytcheva, I., & Bachmann, O. (2014). LA-ICP-MS Pb–U dating of young zircons from the Kos–Nisyros
volcanic centre, SE Aegean arc. Journal of Analytical Atomic Spectrometry, 29, 963–970. https://doi.org/10.1039/C4JA00009A
Gupta, S. (1997). Himalayan drainage patterns and the origin of fluvial megafans in the Ganges foreland basin. Geology, 25(1), 11–14. https://
doi.org/10.1130/0091-7613(1997)025<0011:hdpato>2.3.co;2
Hilgen, F., Lourens, L., Van Dam, J. A., Beu, A., Boyes, A., Cooper, R., et al. (2012). The neogene period. In F. M. Gradstein et al. (Eds.), The
geologic time scale (pp. 923–978). Amsterdam: Elsevier. https://doi.org/10.1016/b978-0-444-59425-9.00029-9
Hirschmiller, J., Grujic, D., Bookhagen, B., Coutand, I., Huyghe, P., Mugnier, J.-L., & Ojha, T. (2014). What controls the growth of the Himalayan
foreland fold-and-thrust belt? Geology, 42(3), 247–250. https://doi.org/10.1130/g35057.1
Horstwood, M. S., Košler, J., Gehrels, G., Jackson, S. E., McLean, N. M., Paton, C., et al. (2016). Community-derived standards for LA-ICP-MS
U-(Th-) Pb geochronology—Uncertainty propagation, age interpretation and data reporting. Geostandards and Geoanalytical Research,
40(3), 311–332. https://doi.org/10.1111/j.1751-908x.2016.00379.x
Hou, Z.-Q., Zheng, Y.-C., Zeng, L.-S., Gao, L.-E., Huang, K.-X., Li, W., et al. (2012). Eocene–Oligocene granitoids in southern Tibet: Constraints on
crustal anatexis and tectonic evolution of the Himalayan orogen. Earth and Planetary Science Letters, 349-350, 38–52. https://doi.org/
10.1016/j.epsl.2012.06.030
Hu, X., Garzanti, E., Wang, J., Huang, W., An, W., & Webb, A. (2016). The timing of India-Asia collision onset—Facts, theories, controversies.
Earth-Science Reviews, 160, 264–299. https://doi.org/10.1016/j.earscirev.2016.07.014
Huang, C., Zhao, Z., Li, G., Zhu, D.-C., Liu, D., & Shi, Q. (2017). Leucogranites in Lhozag, southern Tibet: Implications for the tectonic evolution
of the eastern Himalaya. Lithos, 294-295, 246–262. https://doi.org/10.1016/j.lithos.2017.09.014
Huyghe, P., Galy, A., Mugnier, J.-L., & France-Lanord, C. (2001). Propagation of the thrust system and erosion in the Lesser Himalaya: Geochemical
and sedimentological evidence. Geology, 29(11), 1007–1010. https://doi.org/10.1130/0091-7613(2001)029<1007:pottsa>2.0.co;2
Huyghe, P., Mugnier, J. L., Gajurel, A. P., & Delcaillau, B. (2005). Tectonic and climatic control of the changes in the sedimentary record of the
Karnali River section (Siwaliks of western Nepal). Island Arc, 14(4), 311–327. https://doi.org/10.1111/j.1440-1738.2005.00500.x
Imayama, T., & Arita, K. (2008). Nd isotopic data reveal the material and tectonic nature of the Main central thrust zone in Nepal Himalaya.
Tectonophysics, 451(1–4), 265–281. https://doi.org/10.1016/j.tecto.2007.11.051
Jackson, S. E., Pearson, N. J., Griffin, W. L., & Belousova, E. A. (2004). The application of laser ablation-inductively coupled plasma-mass
spectrometry to in situ U–Pb zircon geochronology. Chemical Geology, 211(1-2), 47–69. https://doi.org/10.1016/j.
chemgeo.2004.06.017
Jacobsen, S. B., & Wasserburg, G. (1980). Sm-Nd isotopic evolution of chondrites. Earth and Planetary Science Letters, 50(1), 139–155. https://
doi.org/10.1016/0012-821x(80)90125-9
Kohn, M. J. (2014). Himalayan metamorphism and its tectonic implications. Annual Review of Earth and Planetary Sciences, 42(1), 381–419.
https://doi.org/10.1146/annurev-earth-060313-055005
Kumar, R. (1993). Coalescence megafan: Multistorey sandstone complex of the late-orogenic (Mio-Pliocene) sub-Himalayan belt, Dehra Dun,
India. Sedimentary Geology, 85(1-4), 327–337. https://doi.org/10.1016/0037-0738(93)90091-i
Kumar, R., Ghosh, S. K., Mazari, R. K., & Sangode, S. J. (2003). Tectonic impact on the fluvial deposits of Plio-Pleistocene Himalayan foreland
basin, India. Sedimentary Geology, 158(3-4), 209–234. https://doi.org/10.1016/s0037-0738(02)00267-1

MANDAL ET AL. 140


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

Kumar, R., Ghosh, S. K., & Sangode, S. J. (1999). Evolution of a Neogene fluvial system in a Himalayan foreland basin, India. In A. Macfarlane, R.
B. Sorkhabi, & J. Quade (Eds.), Himalaya and Tibet: Mountain roots to mountain tops, Geological Society of America Special Papers, (Vol. 328,
pp. 239–256). https://doi.org/10.1130/0-8137-2328-0.239
Kumar, R., & Nanda, A. (1989). Sedimentology of the middle Siwalik subgroup of Mohand area, Dehra Dun valley, India. Geological Society of
India, 34, 597–616.
Kumar, R., Sangode, S. J., & Ghosh, S. K. (2004). A multistorey sandstone complex in the Himalayan Foreland Basin, NW Himalaya, India.
Journal of Asian Earth Sciences, 23(3), 407–426. https://doi.org/10.1016/s1367-9120(03)00176-7
Lavé, J., & Avouac, J.-P. (2000). Active folding of fluvial terraces across the Siwaliks Hills, Himalayas of central Nepal. Journal of Geophysical
Research, 105(B3), 5735–5770. https://doi.org/10.1029/1999JB900292
Le Fort, P. (1975). Himalayas: The collided range. Present knowledge of the continental arc. American Journal of Science, 275, 1–44.
Le Fort, P. (1986). Metamorphism and magmatism during the Himalayan collision. Geological Society, London, Special Publications, 19(1),
159–172. https://doi.org/10.1144/gsl.sp.1986.019.01.08
Liu, Z.-C., Wu, F.-Y., Ding, L., Liu, X.-C., Wang, J.-G., & Ji, W.-Q. (2016). Highly fractionated Late Eocene (~35 Ma) leucogranite in the Xiaru Dome,
Tethyan Himalaya, South Tibet. Lithos, 240, 337–354. https://doi.org/10.1016/j.lithos.2015.11.026
Liu, Z.-C., Wu, F.-Y., Qiu, Z.-L., Wang, J.-G., Liu, X.-C., Ji, W.-Q., & Liu, C.-Z. (2017). Leucogranite geochronological constraints on the termination
of the South Tibetan Detachment in eastern Himalaya. Tectonophysics, 721, 106–122. https://doi.org/10.1016/j.tecto.2017.08.019
Lupker, M., France-Lanord, C., Galy, V., Lavé, J., & Kudrass, H. (2013). Increasing chemical weathering in the Himalayan system since the Last
Glacial Maximum. Earth and Planetary Science Letters, 365, 243–252. https://doi.org/10.1016/j.epsl.2013.01.038
Mandal, S., Robinson, D. M., Khanal, S., & Das, O. (2015). Redefining the tectonostratigraphic and structural architecture of the Almora klippe
and the Ramgarh–Munsiari thrust sheet in NW India. Geological Society, London, Special Publications, 412(1), 247–269. https://doi.org/
10.1144/sp412.6
Mandal, S., Robinson, D. M., Kohn, M. J., Khanal, S., Das, O., & Bose, S. (2016). Zircon U-Pb ages and Hf isotopes of the Askot klippe, Kumaun,
northwest India: Implications for Paleoproterozoic tectonics, basin evolution and associated metallogeny of the northern Indian cratonic
margin. Tectonics, 35, 965–982. https://doi.org/10.1002/2015TC004064
Martin, A. J., DeCelles, P. G., Gehrels, G. E., Patchett, P. J., & Isachsen, C. (2005). Isotopic and structural constraints on the location of the Main
Central thrust in the Annapurna Range, central Nepal Himalaya. Geological Society of America Bulletin, 117(7), 926. https://doi.org/10.1130/
b25646.1
McKenzie, N. R., Hughes, N. C., Myrow, P. M., Xiao, S., & Sharma, M. (2011). Correlation of Precambrian–Cambrian sedimentary successions
across northern India and the utility of isotopic signatures of Himalayan lithotectonic zones. Earth and Planetary Science Letters, 312(3-4),
471–483. https://doi.org/10.1016/j.epsl.2011.10.027
Meigs, A. J., Burbank, D. W., & Beck, R. A. (1995). Middle-late Miocene (>10 Ma) formation of the Main Boundary thrust in the western
Himalaya. Geology, 23(5), 423–426. https://doi.org/10.1130/0091-7613(1995)023<0423:mlmmfo>2.3.co;2
Mercier, J., Braun, J., & van der Beek, P. (2017). Do along-strike tectonic variations in the Nepal Himalaya reflect different stages in the
accretion cycle? Insights from numerical modeling. Earth and Planetary Science Letters, 472, 299–308. https://doi.org/10.1016/j.epsl.2017.
04.041
Métivier, F., Gaudemer, Y., Tapponnier, P., & Klein, M. (1999). Mass accumulation rates in Asia during the Cenozoic. Geophysical Journal
International, 137(2), 280–318. https://doi.org/10.1046/j.1365-246x.1999.00802.x
Miller, C., Thöni, M., Frank, W., Grasemann, B., Klötzli, U., Guntli, P., & Draganits, E. (2001). The early Palaeozoic magmatic event in the
northwest Himalaya, India: Source, tectonic setting and age of emplacement. Geological Magazine,
138(03). https://doi.org/10.1017/s0016756801005283
Mishra, P., & Mukhopadhyay, D. K. (2002). Balanced structural models of Mohand and Santaurgarh ramp anticlines, Himalayan foreland fold-
thrust belt, Dehra Dun re-entrant Uttaranchal. Journal Geological Society of India, 60, 649–661.
Mishra, P., & Mukhopadhyay, D. K. (2012). Structural evolution of the frontal fold–thrust belt, NW Himalayas from sequential restoration of
balanced cross-sections and its hydrocarbon potential. Geological Society, London, Special Publications, 366(1), 201–228. https://doi.org/
10.1144/sp366.6
Molnar, P., Boos, W. R., & Battisti, D. S. (2010). Orographic controls on climate and paleoclimate of Asia: Thermal and mechanical roles for
the Tibetan Plateau. Annual Review of Earth and Planetary Sciences, 38(1), 77–102. https://doi.org/10.1146/annurev-earth-040809-
152456
Morell, K. D., Sandiford, M., Kohn, B., Codilean, A., Fülöp, R.-H., & Ahmad, T. (2017). Current strain accumulation in the hinterland of the
northwest Himalaya constrained by landscape analyses, basin-wide denudation rates, and low temperature thermochronology.
Tectonophysics, 721, 70–89. https://doi.org/10.1016/j.tecto.2017.09.007
Mukherjee, S. (2015). A review on out-of-sequence deformation in the Himalaya. Geological Society, London, Special Publications, 412(1),
67–109. https://doi.org/10.1144/sp412.13
Mukhopadhyay, D. K., & Mishra, P. (2005). A balanced cross section across the Himalayan frontal fold-thrust belt, Subathu area, Himachal
Pradesh, India: Thrust sequence, structural evolution and shortening. Journal of Asian Earth Sciences, 25(5), 735–746. https://doi.org/
10.1016/j.jseaes.2004.07.007
Myrow, P. M., Hughes, N. C., Goodge, J. W., Fanning, C. M., Williams, I. S., Peng, S., et al. (2010). Extraordinary transport and mixing of sediment
across Himalayan central Gondwana during the Cambrian-Ordovician. Geological Society of America Bulletin, 122(9–10), 1660–1670.
https://doi.org/10.1130/b30123.1
Myrow, P. M., Hughes, N. C., Paulsen, T. S., Williams, I. S., Parcha, S. K., Thompson, K. R., et al. (2003). Integrated tectonostratigraphic analysis of
the Himalaya and implications for its tectonic reconstruction. Earth and Planetary Science Letters, 212(3–4), 433–441. https://doi.org/
10.1016/s0012-821x(03)00280-2
Najman, Y. (2006). The detrital record of orogenesis: A review of approaches and techniques used in the Himalayan sedimentary basins.
Earth-Science Reviews, 74, 1–72. https://doi.org/10.1016/j.earscirev.2005.04.004
Najman, Y., Bickle, M., & Chapman, H. (2000). Early Himalayan exhumation: Isotopic constraints from the Indian foreland basin. Terra Nova,
12(1), 28–34. https://doi.org/10.1046/j.1365-3121.2000.00268.x
Najman, Y., Bickle, M., Garzanti, E., Pringle, M., Barfod, D., Brozovic, N., et al. (2009). Reconstructing the exhumation history of the Lesser
Himalaya, NW India, from a multitechnique provenance study of the foreland basin Siwalik Group. Tectonics, 28, TC5018. https://doi.org/
10.1029/2009TC002506
Najman, Y., Bickle, M., Garzanti, E., Pringle, M., Barfod, D., Brozovic, N., et al. (2010). Correction to “Reconstructing the exhumation history of
the Lesser Himalaya, NW India, from a multitechnique provenance study of the foreland basin Siwalik Group”. Tectonics, 29, TC6006.
https://doi.org/10.1029/2010TC002778

MANDAL ET AL. 141


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

Najman, Y., Jenks, D., Godin, L., Boudagher-Fadel, M., Millar, I., Garzanti, E., et al. (2017). The Tethyan Himalayan detrital record shows that
India–Asia terminal collision occurred by 54 Ma in the Western Himalaya. Earth and Planetary Science Letters, 459, 301–310. https://doi.org/
10.1016/j.epsl.2016.11.036
Najman, Y., Johnson, K., White, N., & Oliver, G. (2004). Evolution of the Himalayan foreland basin, NW India. Basin Research, 16(1), 1–24. https://
doi.org/10.1111/j.1365-2117.2004.00223.x
Pagani, M., Zachos, J. C., Freeman, K. H., Tipple, B., & Bohaty, S. (2005). Marked decline in atmospheric carbon dioxide concentrations during
the Paleogene. Science, 309(5734), 600–603. https://doi.org/10.1126/science.1110063
Parrish, R. R., & Hodges, V. (1996). Isotopic constraints on the age and provenance of the Lesser and Greater Himalayan sequences, Nepalese
Himalaya. Geological Society of America Bulletin, 108(7), 904–911. https://doi.org/10.1130/0016-7606(1996)108<0904:icotaa>2.3.co;2
Paton, C., Woodhead, J. D., Hellstrom, J. C., Hergt, J. M., Greig, A., & Maas, R. (2010). Improved laser ablation U-Pb zircon geochronology
through robust downhole fractionation correction. Geochemistry, Geophysics, Geosystems, 11, Q0AA06. https://doi.org/10.1029/
2009GC002618
Powers, P. M., Lillie, R. J., & Yeats, R. S. (1998). Structure and shortening of the Kangra and Dehra Dun reentrants, sub-Himalaya, India.
Geological Society of America Bulletin, 110(8), 1010–1027. https://doi.org/10.1130/0016-7606(1998)110<1010:sasotk>2.3.co;2
Quade, J., & Cerling, T. E. (1995). Expansion of C4 grasses in the Late Miocene of northern Pakistan: Evidence from stable isotopes in paleosols.
Palaeogeography, Palaeoclimatology, Palaeoecology, 115(1-4), 91–116. https://doi.org/10.1016/0031-0182(94)00108-K
Ratschbacher, L., Frisch, W., Liu, G., & Chen, C. (1994). Distributed deformation in southern and western Tibet during and after the India-Asia
collision. Journal of Geophysical Research, 99(B10), 19,917–19,945. https://doi.org/10.1029/94JB00932
Ravikant, V., Wu, F.-Y., & Ji, W.-Q. (2011). U–Pb age and Hf isotopic constraints of detrital zircons from the Himalayan foreland Subathu sub-
basin on the Tertiary palaeogeography of the Himalaya. Earth and Planetary Science Letters, 304(3-4), 356–368. https://doi.org/10.1016/j.
epsl.2011.02.009
Richards, A., Argles, T., Harris, N., Parrish, R., Ahmad, T., Darbyshire, F., & Draganits, E. (2005). Himalayan architecture constrained by isotopic
tracers from clastic sediments. Earth and Planetary Science Letters, 236(3-4), 773–796. https://doi.org/10.1016/j.epsl.2005.05.034
Robinson, D., DeCelles, P., Garzione, C., Pearson, O., Harrison, T., & Catlos, E. (2003). Kinematic model for the Main Central thrust in Nepal.
Geology, 31(4), 359–362. https://doi.org/10.1130/0091-7613(2003)031<0359:kmftmc>2.0.co;2
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 Bulletin, 118, 865–885. https://doi.org/10.1130/b25911
Robinson, D. M., DeCelles, P. G., Patchett, P. J., & Garzione, C. N. (2001). The kinematic evolution of the Nepalese Himalaya interpreted from
Nd isotopes. Earth and Planetary Science Letters, 192(4), 507–521. https://doi.org/10.1016/s0012-821x(01)00451-4
Robinson, D. M., & Martin, A. J. (2014). Reconstructing the Greater Indian margin: A balanced cross section in central Nepal focusing on the
Lesser Himalayan duplex. Tectonics, 33, 2143–2168. https://doi.org/10.1002/2014TC003564
Robinson, D. M., & McQuarrie, N. (2012). Pulsed deformation and variable slip rates within the central Himalayan thrust belt. Lithosphere, 4(5),
449–464. https://doi.org/10.1130/l204.1
Romer, R. L., & Hahne, K. (2010). Life of the Rheic Ocean: Scrolling through the shale record. Gondwana Research, 17, 236–253. https://doi.org/
10.1016/j.gr.2009.09.004
Rubatto, D., Chakraborty, S., & Dasgupta, S. (2013). Timescales of crustal melting in the Higher Himalayan Crystallines (Sikkim, Eastern
Himalaya) inferred from trace element-constrained monazite and zircon chronology. Contributions to Mineralogy and Petrology, 165(2),
349–372. https://doi.org/10.1007/s00410-012-0812-y
Rudnick, R. L., & Gao, S. (2003). Composition of the continental crust. In R. L. Rudnick, H. D. Holland, & K. K. Turekian (Eds.), Treatise on geo-
chemistry (Vol. 3, pp. 1–64). Amsterdam: Elsevier. https://doi.org/10.1016/b978-0-08-095975-7.00301-6
Sangode, S., Kumar, R., & Ghosh, S. K. (1996). Magnetic polarity stratigraphy of the Siwalik sequence of Haripur area (HP), NW Himalaya.
Journal Geological Society of India, 47, 683–704.
Sangode, S., Kumar, R., & Ghosh, S. K. (1999). Palaeomagnetic and rock magnetic perspectives on the post-collision continental sediments of
the Himalaya, India. In T. Radhakrishna, & J. D. A. Piper (Eds.), The Indian subcontinent and Gondwana: A palaeomagnetic and rock magnetic
perspective, Memoir Geological Society of India, (Vol. 44, pp. 221–248).
Sanyal, P., Sarkar, A., Bhattacharya, S. K., Kumar, R., Ghosh, S. K., & Agrawal, S. (2010). Intensification of monsoon, microclimate and asyn-
chronous C4 appearance: Isotopic evidence from the Indian Siwalik sediments. Palaeogeography, Palaeoclimatology, Palaeoecology,
296(1-2), 165–173. https://doi.org/10.1016/j.palaeo.2010.07.003
10
Scherler, D., Bookhagen, B., & Strecker, M. R. (2014). Tectonic control on Be-derived erosion rates in the Garhwal Himalaya, India. Journal of
Geophysical Research: Earth Surface, 119, 83–105. https://doi.org/10.1002/2013JF002955
Searle, M., Noble, S., Hurford, A. J., & Rex, D. (1999). Age of crustal melting, emplacement and exhumation history of the Shivling leucogranite,
Garhwal Himalaya. Geological Magazine, 136(5), 513–525. https://doi.org/10.1017/s0016756899002885
Singh, S., Awasthi, A., Parkash, B., & Kumar, S. (2013). Tectonics or climate: What drove the Miocene global expansion of C4 grasslands?
International Journal of Earth Sciences, 102(7), 2019–2031. https://doi.org/10.1007/s00531-013-0893-5
Singh, S. K., Rai, S. K., & Krishnaswami, S. (2008). Sr and Nd isotopes in river sediments from the Ganga Basin: Sediment provenance and
spatial variability in physical erosion. Journal of Geophysical Research, 113, F03006. https://doi.org/10.1029/2007JF000909
Sláma, J., Koˇsler, J., Condon, D. J., Crowley, J. L., Gerdes, A., Hanchar, J. M., et al. (2008). Pleˇsovice zircon—A new natural reference material
for U–Pb and Hf isotopic microanalysis. Chemical Geology, 249(1-2), 1–35. https://doi.org/10.1016/j.chemgeo.2007.11.005
Spencer, C. J., Harris, R. A., & Dorais, M. J. (2012a). The metamorphism and exhumation of the Himalayan metamorphic core, eastern Garhwal
region, India. Tectonics, 31, TC1007. https://doi.org/10.1029/2010TC002853
Spencer, C. J., Harris, R. A., & Dorais, M. J. (2012b). Depositional provenance of the Himalayan metamorphic core of Garhwal region, India:
Constrained by U–Pb and Hf isotopes in zircons. Gondwana Research, 22(1), 26–35. https://doi.org/10.1016/j.gr.2011.10.004
Srivastava, P., & Mitra, G. (1994). Thrust geometries and deep structure of the outer and lesser Himalaya, Kumaon and Garhwal (India):
Implications for evolution of the Himalayan fold-and-thrust belt. Tectonics, 13(1), 89–109. https://doi.org/10.1029/93TC01130
Stevens, V., & Avouac, J. P. (2016). Millenary Mw> 9.0 earthquakes required by geodetic strain in the Himalaya. Geophysical Research Letters,
43, 1118–1123. https://doi.org/10.1002/2015GL067336
Stübner, K., Grujic, D., Dunkl, I., Thiede, R., & Eugster, P. (2018). Pliocene episodic exhumation and the significance of the Munsiari thrust in the
northwestern Himalaya. Earth and Planetary Science Letters, 481, 273–283. https://doi.org/10.1016/j.epsl.2017.10.036
Tripathi, J. K., Bock, B., & Rajamani, V. (2013). Nd and Sr isotope characteristics of Quaternary Indo-Gangetic plain sediments: Source dis-
tinctiveness in different geographic regions and its geological significance. Chemical Geology, 344, 12–22. https://doi.org/10.1016/j.
chemgeo.2013.02.016
Valdiya, K. (1980). Geology of Kumaun Lesser Himalaya, . Interim Record: Dehradun, Wadia Institute of Himalayan Geology.

MANDAL ET AL. 142


19449194, 2019, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2018TC005200, Wiley Online Library on [04/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2018TC005200

van der Beek, P., Robert, X., Mugnier, J.-L., Bernet, M., Huyghe, P., & Labrin, E. (2006). Late Miocene—Recent exhumation of the central
Himalaya and recycling in the foreland basin assessed by apatite fission-track thermochronology of Siwalik sediments, Nepal. Basin
Research, 18(4), 413–434. https://doi.org/10.1111/j.1365-2117.2006.00305.x
Vannay, J. C., Grasemann, B., Rahn, M., Frank, W., Carter, A., Baudraz, V., & Cosca, M. (2004). Miocene to Holocene exhumation of metamorphic
crustal wedges in the NW Himalaya: Evidence for tectonic extrusion coupled to fluvial erosion. Tectonics, 23, TC1014. https://doi.org/
10.1029/2002TC001429
Vögeli, N., Najman, Y., van der Beek, P., Huyghe, P., Wynn, P. M., Govin, G., et al. (2017). Lateral variations in vegetation in the Himalaya since the
Miocene and implications for climate evolution. Earth and Planetary Science Letters, 471, 1–9. https://doi.org/10.1016/j.epsl.2017.04.037
Webb, A. A. G. (2013). Preliminary balanced palinspastic reconstruction of Cenozoic deformation across the Himachal Himalaya (north-
western India). Geosphere, 9(3), 572–587. https://doi.org/10.1130/GES00787.1
Webb, A. A. G., Yin, A., Harrison, T. M., Célérier, J., Gehrels, G. E., Manning, C. E., & Grove, M. (2011). Cenozoic tectonic history of the Himachal
Himalaya (northwestern India) and its constraints on the formation mechanism of the Himalayan orogen. Geosphere, 7(4), 1013–1061.
https://doi.org/10.1130/GES00627.1
Weinberg, R. (2016). Himalayan leucogranites and migmatites: Nature, timing and duration of anatexis. Journal of Metamorphic Geology,
34(8), 821–843. https://doi.org/10.1111/jmg.12204
White, N. M., Pringle, M., Garzanti, E., Bickle, M., Najman, Y., Chapman, H., & Friend, P. (2002). Constraints on the exhumation and erosion of
the High Himalayan Slab, NW India, from foreland basin deposits. Earth and Planetary Science Letters, 195(1-2), 29–44. https://doi.org/
10.1016/s0012-821x(01)00565-9
Wiedenbeck, M., Alle, P., Corfu, F., Griffin, W., Meier, M., Oberli, F. V., et al. (1995). Three natural zircon standards for U-Th-Pb, Lu-Hf,
trace element and REE analyses. Geostandards and Geoanalytical Research, 19(1), 1–23. https://doi.org/10.1111/j.1751-908X.1995.
tb00147.x
Yin, A. (2006). Cenozoic tectonic evolution of the Himalayan orogen as constrained by along-strike variation of structural geometry, exhu-
mation history, and foreland sedimentation. Earth-Science Reviews, 76(1-2), 1–131. https://doi.org/10.1016/j.earscirev.2005.05.004
Yu, H., Webb, A. A. G., & He, D. (2015). Extrusion vs. duplexing models of Himalayan mountain building 1: Discovery of the Pabbar thrust
confirms duplex-dominated growth of the northwestern Indian Himalaya since Mid-Miocene. Tectonics, 34, 313–333. https://doi.org/
10.1002/2014TC003589
Zeng, L., Gao, L.-E., Xie, K., & Liu-Zeng, J. (2011). Mid-Eocene high Sr/Y granites in the Northern Himalayan Gneiss Domes: Melting thickened
lower continental crust. Earth and Planetary Science Letters, 303(3-4), 251–266. https://doi.org/10.1016/j.epsl.2011.01.005
Zheng, Y., Hou, Z., Fu, Q., Zhu, D.-C., Liang, W., & Xu, P. (2016). Mantle inputs to Himalayan anatexis: Insights from petrogenesis of the Miocene
Langkazi leucogranite and its dioritic enclaves. Lithos, 264, 125–140. https://doi.org/10.1016/j.lithos.2016.08.019

MANDAL ET AL. 143

You might also like