You are on page 1of 50

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/319406901

Three-Dimensional Crustal Architecture Beneath the Sikkim Himalaya and Its


Relationship to Active Deformation

Article  in  Journal of Geophysical Research: Solid Earth · November 2017


DOI: 10.1002/2017jb014506

CITATIONS READS

6 468

2 authors:

Himangshu Paul Supriyo Mitra


National Geophysical Research Institute Indian Institute of Science Education and Research Kolkata
11 PUBLICATIONS   87 CITATIONS    64 PUBLICATIONS   958 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

High frequecy seismic attenuation in Himalaya View project

Study of seismic anisotropy in Himalayas View project

All content following this page was uploaded by Himangshu Paul on 09 October 2017.

The user has requested enhancement of the downloaded file.


Three-Dimensional crustal architecture beneath the
Sikkim Himalaya and its relationship to active
deformation
1 1
Himangshu Paul and Supriyo Mitra

Himangshu Paul, heman2007s@gmail.com

1
Department of Earth Sciences, Indian

Institute of Science Education and Research

Kolkata, Mohanpur-741246, West Bengal,

India.

This article has been accepted for publication and undergone full peer review but has not been through
the copyediting, typesetting, pagination and proofreading process, which may lead to differences
between this version and the Version of Record. Please cite this article as doi: 10.1002/2017JB014506

2017
c American Geophysical Union. All Rights Reserved.
Abstract. We study the 3-D variation of the crustal structure of the Sikkim

Himalaya using broadband seismological data acquired from a focussed net-

work of seven stations spanning the Lesser, Higher and Tethyan Himalaya.

Common conversion point stacking of receiver functions recorded along an

across-strike profile of the Himalaya reveals first order northward dip on the

Main Himalayan Thrust (MHT), a mid-crustal discontinuity and the Moho,

along with higher order lateral variations. 3-D images generated from joint

inversion of receiver functions and surface wave dispersions show that the

MHT has a ramp-flat-ramp geometry. The ramps are located beneath the

Lesser Himalaya and the Tethyan Himalaya with dips of ∼7◦ and ∼15◦ re-

spectively, connected by flat segments. The ramp beneath the Lesser Himalaya

forms a dome structure, up-warping the thrust sheets associated with the

Peling and Main Central Thrust. The erosional surface of this dome forms

the arcuate geometry of thrusts observed in the Lesser Himalaya. The thick-

ness of the underthrust Indian crust is 35–42 km, and has an average VS of

3.63 km/s, similar to that of the Indian Shield crust. The Moho also has dome-

like structures separated by elongated, deeper sections trending NW-SE. These

are intersected by steeply-dipping transverse low-velocity zones, oblique to

the strike of the Himalaya. We conjecture that these low-velocity zones are

the dextral-strike slip faults known to be active beneath the Sikkim Himalaya.

The observed alternate shallow and deep segments of the Moho must be a

consequence of several cycles of strike-slip displacement on these transverse

faults.

2017
c American Geophysical Union. All Rights Reserved.
Keypoints:

• Importance of 3-D deformation of the crust in continent-continent col-

lision regime

• Clear observation of the ramp-flat-ramp geometry of the MHT

• Observation of regional-scale lateral heterogeneity within the crust

• Role of transverse tectonics in shaping observed structure

2017
c American Geophysical Union. All Rights Reserved.
1. Introduction

Sikkim is an Indian state bordering Nepal, Tibet and Bhutan (Fig. 1a). The Sikkim

Himalaya lies at the junction of multiple modes of deformation [England and Houseman,

1986; England and Molnar , 1997; Yin, 2000; Yin and Harrison, 2000]. To the west,

the Central Himalaya deforms by arc-normal crustal shortening [Avouac and Tapponnier ,

1993; Avouac, 2003], while to the east, northeast India undergoes oblique convergence

across the Eastern Himalaya, and E–W convergence across the Indo–Burma subduction

zone [Ni et al., 1989; Mitra et al., 2005; Kumar et al., 2015; Li et al., 2008; Lei et al.,

2014; Lei and Zhao, 2016], resulting in widespread deformation south of the Himalaya.

The whole thickness of the Sikkim Himalayan crust is seismogenic, displaying two distinct

regimes of tectonic deformation: (a) crustal shortening within the Himalayan wedge, and

(b) strike-slip motion within the underthrust Indian crust [Paul et al., 2015]. Available 1-

D shear-wave velocity models from previous seismological studies in the Sikkim Himalaya

have highlighted first order variation in crustal structure [Singh et al., 2010; Acton et al.,

2011]. However, detailed 3-D imaging of the major subsurface discontinuities (viz. the

Main Himalayan Thrust (MHT), the Moho and the mid-crustal discontinuity) and their

relationship to the active deformation has not been attempted, motivating our broadband

seismological imaging of the crust and upper mantle structure of the Sikkim Himalaya.

Additionally, the lower Main Central Thrust (MCT) and the higher Peling Thrust (PT)

within the Sikkim Himalaya form an arcuate geometry on the surface [Bhattacharyya and

Mitra, 2009] (Fig 1a). This contrasts with adjacent segments of the Himalaya where

these thrusts run parallel to the Himalayan arc, the Main Boundary Thrust and the

2017
c American Geophysical Union. All Rights Reserved.
Himalayan Frontal Thrust [Yin, 2006]. Correlation of the structural variations observed

at the surface with crustal structure at depth is a prerequisite for kinematic analysis of

active deformation, and has never been attempted for the Sikkim Himalaya.

The first comprehensive image of the underthrusting Indian plate beneath the Sikkim–

Darjeeling Himalaya was provided by Acton et al. [2011] using receiver function analysis.

These results were from an approximately N–S profile averaging a broad region across the

Indian foreland basin (western Bengal Basin) and the Sikkim–Darjeeling Himalaya. That

study highlighted that the Moho dips gently to the north with local southward dip beneath

Gangtok in the Lesser Himalaya. The MHT was seen to be dipping gently beneath the

Lesser Himalaya, but its geometry further north was obscured due to lack of high signal-to-

noise data. Singh et al. [2010] used receiver function analysis to highlight the north-dipping

Moho beneath Sikkim with the presence of a ”Moho-doublet” at ∼40 km depth beneath

the Lesser Himalaya. This was interpreted as eclogitization of the granulitic Indian lower

crust, but the authors did not discuss the pressure-temperature regime necessary to cause

eclogitization at such shallow depth. Their study also explored the presence of a strong

layer of anisotropy within a low velocity layer at 20–30 km depth. The starting models for

inversion of receiver functions in both these studies [Singh et al., 2010; Acton et al., 2011]

were based on the 1-D density models proposed from an earlier study of gravity anomalies

across the Sikkim Himalaya [Tiwari et al., 2006]. Magnetotelluric studies by Patro and

Harinarayana [2009] in Sikkim reported the presence of a conductive feature at a depth

of 3–15 km. This was interpreted as molasse sediment from the Siwaliks and Lesser

Himalaya, or as trapped metamorphic fluids lubricating the underthrusting of the Indian

plate. Paul et al. [2015] studied the source mechanisms of the 2011 Sikkim earthquake

2017
c American Geophysical Union. All Rights Reserved.
and its aftershocks, which originated in the underthrust Indian crust, demonstrating the

presence of active transverse strike-slip faults beneath the Sikkim Himalaya. The 2006

Phodong earthquake, which originated within the Himalayan wedge, had a thrust fault

mechanism highlighting the different deformation regimes in the underthrust Indian crust

and the overriding Himalayan wedge.

These studies have indicated that the structure of the underthrust Indian crust beneath

the Sikkim Himalaya plays an important role in partitioning along-strike convergence

across the Eastern Himalaya. Additionally, the configuration of the Himalayan thrusts at

depth and the geometry of the detachment surface (MHT) play a significant role in strain

accumulation within the Himalayan wedge, controlling the rupture size and geometry of

megathrust earthquakes. Through a number of studies, we have developed a fairly con-

sistent image of the large-scale long wavelength variation of the crustal structure beneath

the Eastern Himalaya [Mitra et al., 2005], the Sikkim Himalaya [Acton et al., 2011], the

Nepal Himalaya [Schulte-Pelkum et al., 2005; Nábělek et al., 2009], the Garhwal Himalaya

[Caldwell et al., 2013] and the NW Himalaya [Rai et al., 2006]. However, detailed subsur-

face architecture is still uncertain. The relationships between geological features observed

on the surface and structures at depth are poorly understood. The detailed geometry of

the MHT, the presence/absence of ramp-flat-ramp geometry and the role of the Lesser

Himalayan Duplex in accommodating active deformation beneath the Sikkim Himalaya

are yet to be established. In this study, we use receiver function analysis on seismologi-

cal data acquired from a focussed experiment of seven broadband stations spanning the

Lesser, Higher and Tethyan Himalaya. Our study provides a comprehensive shear-wave

velocity structure of the Sikkim Himalaya and 3-D geometry of the MHT and the Moho.

2017
c American Geophysical Union. All Rights Reserved.
Finally, we examine the relationships between (i) the mapped geological features and the

3-D crustal structure, and (ii) the 3-D crustal structure and the active faults within the

underthrust Indian crust.

2. Data and Methodology

The seismological field experiment in the Sikkim-Darjeeling Himalaya was done in two

phases the first in 2005, and the second in 2008–13. The first phase of the experiment

was operated and funded through a joint project between University of Cambridge and

Indian Institute of Technology (IIT) Kharagpur. The second phase was initiated from

IIT Kharagpur with funding from Department of Science and Technology (DST), India,

and then jointly operated through the UK-India Education and Research Initiative (UK-

IERI) thematic partnership between IISER Kolkata and the University of Cambridge.

Stations Gangtok (GTOK), Chungthang (CHTG) and Lachung (LCHG) were common in

both phases. The other stations Namchi (NAMC), Rabangla (RABN), Mangan (MANG)

and Yumthang (YUMT) were installed in 2008. The stations were deployed along an

approximate N–S profile across Sikkim (Fig. 1a), extending from the Lesser Himalaya to

the Tethyan Himalaya. NAMC and RABN were located in the Lesser Himalaya; GTOK,

MANG, CHTG and LCHG in the Higher Himalaya; YUMT in the Tethyan Himalaya.

GTOK was equipped with a Guralp CMG-3TD seismometer with a flat velocity response

between 120 s and 50 Hz. All the other stations had Guralp CMG-3ESPCD instruments,

with flat velocity response between 60 s and 100 Hz. All stations were equipped with

CMG-DCM/EAM data loggers that recorded data continuously at 100 samples per second.

The data were time stamped using GPS synchronization.

2017
c American Geophysical Union. All Rights Reserved.
2.1. Receiver Function Analysis

Receiver functions are the response of the near receiver structure to incoming teleseis-

mic body waves and are made up of conversions and reverberations from discontinuities

beneath the receiver site. A total of 1430 teleseismic earthquakes (mb ≥ 5.5) from an epi-

central distance range of 30◦ to 95◦ were recorded by the seven broadband stations (Fig.

1b). Waveforms from these events have been used to compute P-wave receiver functions.

First the horizontal components (North-South and East-West) were rotated into radial

and tangential directions to isolate the P-SV energy. This was followed by deconvolution

of the vertical component from the radial and tangential components to remove source

effects and common path effects between the source and the discontinuity. The decon-

volved waveforms are called the radial and the tangential receiver functions. For a 1-D

isotropic velocity structure, the radial receiver function contains all the P-to-S converted

and reverberating energy, while the tangential receiver function has no theoretical energy.

However, for the real Earth, the tangential component has energy and its amplitude is

used to assess the degree of deviation from the 1-D isotropic assumption. We use the

iterative time domain deconvolution technique of Ligorria and Ammon [1999] to compute

the radial and tangential receiver functions. In this technique, the misfit between the

radial component and the convolution of the vertical component with an iteratively up-

dated spike train is minimized in the least-squares sense. Iterations are continued upto a

fixed number (i.e. 200) or until the misfit value falls below a threshold (set at 0.005). To

eliminate high frequency noise, a low-pass Gaussian filter of width 2.5 (corner frequency

∼1.2 Hz) is applied to the waveform. The receiver functions obtained through this time

domain iterative technique are assessed for quality by the percentage fit to the observed

2017
c American Geophysical Union. All Rights Reserved.
radial waveform. For our analysis we selected only those waveforms which had a fit of

≥80%. Using this criterion, a total of 1458 receiver functions were shortlisted for our

study.

In order to enhance signal-to-noise ratio and increase coherence in the signal, receiver

functions were stacked in the following ways: spatial stack, common conversion point

(CCP) stack (Fig. 2), and depth (H) vs VP /VS (K) stack. Spatial stacking is performed

by averaging receiver functions in spatial bins using their piercing points (Fig. 2a) on

a given discontinuity (Moho, in our case). This has the advantage of highlighting the

coherent features in the structure, irrespective of the station at which the waveform is

recorded. In spatial stacking, all receiver functions are projected onto a selected profile

and binned by distance along the profile (Fig. 2b). Piercing points have been calculated

using the crustal velocity structure from Acton et al. [2011] across the Sikkim Himalaya.

The crustal thickness beneath the Himalaya is larger than that in the global 1-D velocity

models, so using this velocity structure has the advantage of computing Moho conversion

depths using crustal rather than upper mantle velocities. The spatial plot is equivalent to

a time section with the y-axis representing time offset of the receiver function from the Pp

arrival. The spatial plot has the advantage of aligning coherent conversions to highlight

discontinuities, but does not represent the true geometry of the discontinuity.

CCP stacking uses similar piercing point information to spatial stacking, but con-

verts the arrival to depth beneath the surface (i.e. depth section). Additionally, in the

CCP stack the amplitude of the P-to-S converted phase is colour-coded to highlight the

impedance contrast across the discontinuity (Fig. 2c). For the CCP stack, we employed

the technique of Zhu [2000] where the receiver functions are back-traced using ray theory

2017
c American Geophysical Union. All Rights Reserved.
through the velocity model to their piercing-point location. The receiver function ampli-

tude values at each point are corrected for incidence angle effect and projected onto the

ray paths according to the velocity model. The volume along the profile is then divided

into bins of a certain size and all the amplitudes within each bin are averaged. The bin

size of our model was 5 km wide, 1 km long and 0.6 km high. For a maximum frequency

of 1.2 Hz, the vertical resolution is estimated to be 1.6 km. Note that both these stacked

images (spatial and CCP) are 2-D profiles and suppress information on the lateral varia-

tion of the subsurface discontinuities. Lateral variations imaged through 3-D mapping of

the discontinuities will be discussed in Section 2.2.

In H-K stacking, time domain waveforms are transformed into depth domain to esti-

mate the thickness (H) and average VP /VS (K) of the crustal layer. The arrival times

of P-to-S converted and multiple reverberating phases in a radial receiver function are

dependent on layer thickness (H) and average VP /VS , for a given VP [Zhu and Kanamori ,

2000]. H-K stacking uses a grid search over a range of H and K values and calculates

the sum s(H, K) of the amplitudes of the phases (conversion and reverberations) at each

grid point. A contour map of s(H, K) is maximized when all the three phases are stacked

coherently and approximate values of H and K are obtained. The effects of lateral vari-

ation in the subsurface structure can be highlighted by performing the H-K stack on

receiver functions clustered in narrow bins of back-azimuth and distance. In our anal-

ysis, receiver functions from each station were divided into clusters based on a narrow

range of back-azimuth and ray parameter. From all the stations, a total of 39 clusters

were formed with significant number of receiver functions in each of them. The H-K

stack for each cluster has been estimated separately for both the entire crust (Moho), and

2017
c American Geophysical Union. All Rights Reserved.
the upper crust (mid-crustal discontinuity). Fig. 3 shows the H-K stack for the Moho

and mid-crustal layer for receiver functions recorded at RABN in the back-azimuth range

127◦ –157◦ , along with the synthetic receiver functions generated by forward modelling of

these two discontinuities. We follow the same method to compute H-K stacks for all the

clusters at every station. The results are summarized in Table 1.

2.2. Joint Inversion for shear-wave velocity structure and 3-D mapping of

discontinuities
Receiver function stacks from every cluster beneath each station are jointly inverted

with Rayleigh-wave group velocity dispersion data to constrain the shear-wave veloc-

ity structure of the crust and uppermost mantle. Receiver functions are sensitive to

impedance-contrast boundaries, while surface wave dispersion data are sensitive to verti-

cal averages of shear-wave velocity beneath the seismograph site. Joint inversion of the

two datasets complement each other and provide optimal shear-wave velocity structure.

Rayleigh-wave group velocity dispersion data, corresponding to the station locations in

Sikkim, were extracted from regional surface wave tomography results of India and Eura-

sia [Gilligan, 2014]. These dispersion curves provide group velocities at discrete periods

between 5 and 70 s. For the joint inversion, we use the linearized least-squares inversion

algorithm of Herrmann and Ammon [2004]. In this method, an a-priori value controls the

influence of each dataset by trading off between the two. A forward model is estimated

based on the values of H and VP /VS from the H-K stacking results. This forward model

is taken as the initial model for the inversion and parameterized by thin layers of fixed

thickness and velocity. The final model, with a minimum number of interfaces, which

fits both the datasets is obtained through iterative least squares inversion. The a-priori

2017
c American Geophysical Union. All Rights Reserved.
weighting between the receiver function and surface wave dispersion data is set to 0.8 and

0.2, respectively. The final model is then coarsely parameterized by combining layers with

similar velocities to obtain a minimum layer model. Fit to the observed receiver function

and dispersion data, for both the models (minimum interfaces and minimum layers) are

compared to highlight the main features in the crust and uppermost mantle structure.

Fig. 4 summarizes the joint inversion results for RABN in the back-azimuth range of

127◦ –157◦ . The main features of the crustal structure are the thin (2–4 km) low-velocity-

layer (LVL) in the shallow crust and the two major positive impedance contrasts, in the

mid-crust and at the Moho. We perform uncertainty analysis on each of these features

and obtain a depth uncertainty of ± 2 km and velocity (VS ) uncertainty of ±0.05 km/s.

To observe the lateral variation of the major discontinuities (the MHT and Moho) be-

neath the Sikkim Himalaya, we devised a method to obtain the map co-ordinates of a given

discontinuity and its spatial depth variation. The depth of the discontinuities (the MHT

and Moho) and the associated VS for a particular cluster is obtained by joint inversion of

receiver functions and Rayleigh-wave group velocity dispersion data. The piercing point

for each receiver function is calculated for the inferred depth (of the discontinuity) using

the obtained velocity model. The piercing point gives the spatial location of the disconti-

nuity at a given depth. This procedure is followed for all receiver functions which provides

a distribution of points in 3-dimensions, representing the depth to a given discontinuity.

We fit a surface through these points using bi-cubic interpolation to obtain a lateral

variation map of the subsurface discontinuity. The bi-cubic interpolation method fits a

continuous surface on a unit square with continuous derivatives. By patching together

such continuous bi-cubic surfaces with continuous derivatives across the boundaries, we

2017
c American Geophysical Union. All Rights Reserved.
obtain a smoothed discontinuity surface. This has the advantage of lateral correlation of

the discontinuity across the region, but can underestimate steep slopes or abrupt change

in depth to the discontinuity. Using this method, we construct 3-D maps of the MHT and

the Moho beneath the Sikkim Himalaya.

3. Results:

3.1. 2-D profiles: Spatial and CCP Stacking results

Receiver functions from all stations were spatially stacked along the AA’ profile in

1 km bins (Fig. 2a,b). The profile is aligned with the station locations to maximize the

region covered by the receiver functions. Positive amplitudes are coloured black while the

negative amplitudes are grey. The P s conversion from the MHT, mid-crust and Moho

are marked by magenta, green and yellow dashed lines, respectively. The shallow low

velocity layer, with a negative arrival, is identified as the MHT. This appears at a time

offset of ∼1 s below the southern end of the profile (NAMC) and with the offset gradually

increasing northward to ∼3 s (YUMT), indicates the northward dip of the structure.

We observe three significant steps on the MHT at distances of ∼35 km, 80–90 km and

∼120 km, along the profile. These steps indicate a sharp change in northward dip on

the MHT. The mid-crustal P s is at a time offset of ∼3.5 s for the first 40 km along

the profile. This offset gradually increases to ∼5 s at 60 km along the profile. At this

point, a sharp decrease in the arrival time of the mid-crustal P s is observed (from ∼4 s

to ∼3 s), which possibly corresponds to a local southward dip of the mid-crust. Further

north, the arrival time from the mid-crust increases steadily northwards from ∼3 s to

∼5 s. The Moho P s phase shows similar characteristics to the mid-crustal phase. At the

southern end of the profile, the offset time of the Moho P s phase is ∼6.5 s, increasing to

2017
c American Geophysical Union. All Rights Reserved.
∼7.5 s at 45 km along the profile. This is followed by a sharp decrease in the offset time,

indicating Moho imbrication or local southward dip. Beyond a distance of 60 km along

the profile, the Moho P s again increases gradually to ∼7 s. Several more arrivals are

also observed in addition to the ones discussed above. Two distinct positive phases are

observed earlier than the MHT arrival at distances of 80–135 km and 115–135 km, along

the profile. These mark discontinuities within the Himalayan wedge, possibly associated

with thrust imbrication underneath the Higher Himalaya. Another positive arrival is

observed immediately following the MHT arrival. This is prominently visible at distances

of 75–135 km along the profile, and becomes less distinguishable elsewhere. This phase

appears to track the MHT and deepens northward. This could be the uppermost crustal

layer of the underthrust Indian plate.

The CCP stack has been constructed along the same profile (AA’, Fig. 2a) and provides

a depth section across the Sikkim Himalaya (Fig. 2c). Impedance contrast discontinuities

with positive gradient are coloured red and those with negative gradient are coloured

blue. The intensity of the colour is scaled by the ratio of the phase amplitude to the

direct arrival P p. The MHT corresponds to the negative impedance contrast arrivals seen

in the shallower part of the crust [Caldwell et al., 2013] while the mid-crust and Moho

are positive impedance contrast arrivals. The MHT, the mid-crustal interface and the

Moho are marked by the same colours as for the spatial plot discussed above. To first

order, the MHT, mid-crust and Moho are all observed to dip to the north. The MHT

has two significant changes in dip, one beneath GTOK and the other beyond LCHG.

We ascribe these to the presence of ramps on the MHT connected by flat sections. The

Moho displays second order dip variations at a number of places, the strongest among

2017
c American Geophysical Union. All Rights Reserved.
these is the region underneath RABN. Here, we observe an apparent Moho-doublet, pos-

sibly due to projection of a 3-dimensional Moho onto a 2-D profile. This feature will be

discussed further in Section 3.3.1. The Moho is also observed to dip locally southward

beneath GTOK and CHTG. This could be an effect of the flexure of the seismogenic

crust resulting in imbrication of the Moho, which has been observed beneath Nepal Hi-

malaya [Hirn et al., 1984]. The mid-crustal boundary displays a similar geometry to the

Moho. Southward dip and imbrication of this boundary are observed beneath GTOK

and CHTG. The lower crustal layer, between the mid-crustal discontinuity and the Moho,

appears to have similar thickness across the Sikkim Himalaya. In addition to these first

order near-horizontal crustal discontinuities, a number of steeply dipping (∼65-70◦ ) low

velocity structures are observed on the CCP at distances between 55 km and 75 km along

the profile (beneath MANG and CHTG). These structures are marked by dashed lines

within an ellipse (Fig. 2c) and are observed to intersect the Moho and the mid-crustal

discontinuity. These low-velocity zones may be active transverse structures within the

underthrust Indian crust, similar to the ones which caused dextral strike slip earthquakes

beneath the Sikkim Himalaya [Paul et al., 2015]. A couple of strong negative arrivals are

observed in the CCP, between stations NAMC and RABN at a depth of ∼20 km, and

between stations RABN and GTOK at depth range 30-40 km. Through forward modeling

of the RFs, we have ascertained that these are multiples from the MHT negative arrival.

3.2. H-K stacking results

The Lesser Himalayan crust is sampled by NAMC and RABN. H-K stacks show the

presence of lateral variation in the depth to the Moho and to the mid-crustal discontinuity

beneath both of these stations. The Moho depth beneath NAMC and RABN varies from

2017
c American Geophysical Union. All Rights Reserved.
43 km to 52 km, with deeper segments at the N-NW and SE back-azimuths and the

shallowest segment in-between (Fig. 5a and Table 1). The mid-crustal depth beneath

the Lesser Himalaya varies between 27 km and 37 km, with E-W lateral variation. The

shallowest segment lies in the middle with deeper segments to the east and to the west

(Fig. 5b). The VP /VS ratio beneath the Lesser Himalaya varies between 1.67 and 1.77

and increases from the east to the west. The mean values of VP /VS for the upper crust

and the whole crust are ∼1.70 and ∼1.72, respectively.

Beneath GTOK in the Higher Himalaya, the Moho depth varies between 50 km and

58 km (west to east), and has a gentle southward dip along the SE back-azimuth (Fig. 5a).

The mid-crustal layer has a shallow segment at ∼28 km in the north, and deepens to

∼36 km to the south and the east (Fig. 5b). Significantly low crustal VP /VS is observed

just west of GTOK (Fig. 5c, d). Further north within the Higher Himalaya, MANG,

CHTG and LCHG shows a gradual northward increase in crustal thickness without any

significant lateral heterogeneity. For MANG, the depth to the mid-crustal discontinuity

and the Moho increases northward from 35 km to 46 km, and 53 km to 62 km, respectively.

The average VP /VS for the upper crust and the whole crust beneath both GTOK and

MANG are found to be ∼1.71 and ∼1.73, respectively. The Moho depth beneath CHTG

increases northward from 55 km to 59 km, but has a southward dip to the south of the

station. The mid-crustal depth also increases northward from 33 to 47 km. The Moho

depth beneath LCHG varies from 57 km to 61 km, increasing to the north and the east.

The depth of the mid-crustal layer is observed to vary from 44 km to 47 km (Fig. 5a, b).

The average VP /VS for both the upper crust and whole crust beneath both CHTG and

LCHG is ∼1.71 (Fig. 5c, d).

2017
c American Geophysical Union. All Rights Reserved.
The Tethyan Himalaya, sampled at YUMT, shows a Moho depth variation from 58 km

to 60 km, with a shallow isolated segment of 53 km in the NE back-azimuth (Fig. 5a).

The mid-crustal layer uniformly dips to the north from 42 km to 47 km. The VP /VS for

the upper crust varies between 1.69 and 1.72, with an average value of ∼1.70. The average

VP /VS for the whole crust is ∼1.73. This reveals that the average VP /VS for the Indian

upper crust is slightly lower than that of the whole crust (Fig. 5c & d) indicating that

the underthrust lower Indian crust is more mafic in composition.

3.3. Joint inversion results

Joint inversion of receiver functions and Rayleigh-wave group velocity dispersion data

has been used to constrain the shear-wave velocity structure beneath the Sikkim Himalaya.

The starting model for the joint inversion is built from the H-K result. The joint inverted

shear-wave velocity model for RABN in the back-azimuth range of 127◦ to 157◦ is plot-

ted in Fig. 4 as an example. We discuss the joint inversion results region-wise for the

Sikkim Himalaya and provide a station-wise summary in Table 2. We have overlaid the

shear-wave velocity models, obtained by joint inversion, onto the CCP stack (Fig. 6a, b)

along the profile AA’ to compare and contrast the results. The major crustal discontinu-

ities, splay faults, and the inferred transverse structures beneath the Lesser, Higher and

Tethyan Himalaya with the average VS listed for each layer, are highlighted in Fig. 6c. A

comparison of the Moho depths obtained from H-K stacking and joint inversion results

reveal a good agreement (Fig. 6d), except for the southern back-azimuth in RABN and

northern back-azimuth in MANG. The differences could be due to deviation from the as-

sumed VP of 6.4 km/s in the H-K analysis. Such deviations could explain the anomalous

VP /VS values in these regions (Fig. 5c).

2017
c American Geophysical Union. All Rights Reserved.
3.3.1. Lesser Himalaya

NAMC and RABN are situated on the low-grade metamorphic rocks of the Lesser

Himalaya. Joint inversion results highlight the presence of lateral heterogeneity beneath

both of these stations (Table 2 and Fig. 6). The MHT is demarcated by a LVL at

shallow depth. South of NAMC, it is observed at a depth of ∼8 km, increasing to ∼10 km

northward, between NAMC and RABN. Further to the north, beyond RABN, the LVL’s

depth reduces to ∼7 km (Table 2). The deeper segment of the MHT is also observed to

the east of both NAMC and RABN and attests to the 3-D configuration of this boundary.

This will be discussed in detail later (Section 3.4.1 and Fig. 7). In the 2-D representation

(Fig. 6), this lateral heterogeneity is masked because of averaging effects, and the MHT

appears as a flat beneath the Lesser Himalaya (depth ∼8 km). It is followed by a ramp

(localized dip ∼10◦ ) south of GTOK (10–12 km). In several places across the Lesser

Himalaya, the MHT appear as a thick (4–6 km) LVL, possibly representing a combination

of the MHT and shallower imbricate faults. Beneath NAMC and RABN, the average VS of

the LVL is modeled to be ∼3.15 km/s and that of the Himalayan wedge (above the MHT)

to be ∼3.3 km/s. The Moho and the mid-crustal layer also reveal large lateral variation

in depth across the region. Shallower segments of the Moho (∼45 km) are observed to the

west of both NAMC and RABN. To their east, the Moho deepens to reach a maximum

depth of ∼52 km, and then again shallows to ∼45 km. These structures dip approximately

E-W and appear as ridges in 3-D. These structures could be due to imbrication of the

crust (discussed in detail in Section 3.4.2), and are clearly observed in the 3-D plot of the

Moho (Fig. 8) but appear as a Moho-doublet in 2-D representation (Fig. 6). The mid-

crustal layer has deeper section (∼33 km) to the west of both NAMC and RABN, shallows

2017
c American Geophysical Union. All Rights Reserved.
to ∼25 km eastwards and again deepens to ∼28 km further east. A similar observation

was also made from the H-K stacking results with comparable depth values (Fig. 5b).

However, this lateral variation is not apparent in the 2-D representation (Fig. 6) due to

averaging of receiver function amplitudes. The average VS for the upper crust and the

lower crust beneath the Lesser Himalaya are modeled to be ∼3.42 km/s and ∼3.82 km/s,

respectively, indicating that the lower crust is more mafic in composition than the upper

crust. The average VS for the whole crust is ∼3.63 km/s. The VS for the uppermost

mantle beneath the Lesser Himalaya is ∼4.22 km/s.

3.3.2. Higher Himalaya

Receiver functions from GTOK, MANG, CHTG and LCHG sample the crystalline com-

plex and high-grade metamorphic rocks of the Higher Himalaya. The lateral variation in

structure observed beneath the Lesser Himalaya continues further north. Beneath GTOK,

in the NW back-azimuth, the MHT is at a depth of ∼7 km. This is surrounded by deeper

segments (10–12 km) eastward and southward, resulting in an east dipping MHT west of

the station (Fig. 7). This feature was not observed in the 2-D image due to averaging

effects (Fig. 6). North of GTOK, the structural variation is mainly north-south. The

MHT depth beneath MANG varies from ∼10 km in the south to ∼13 km in the north,

with a dip of ∼8◦ . Beneath CHTG, the MHT exhibits a southward dip as it shallows

from ∼15 km in the south to ∼13 km in the north. The MHT beneath LCHG is at a

depth of ∼16 km. The average VS of the LVL beneath GTOK, MANG and CHTG is

∼3.23 km/s, and increases marginally beneath LCHG to ∼3.28 km/s. The average VS of

the Himalayan wedge is modeled to be ∼3.29 km/s. From south to north, the average VS

of the LVL progressively approaches that of the Himalayan wedge.

2017
c American Geophysical Union. All Rights Reserved.
The lateral variation in the Moho depth beneath GTOK is highlighted by the presence

of a shallow Moho (∼45 km) beneath east of GTOK, while in all other directions the depth

is ∼54 km. Comparing with Moho depth beneath RABN, we observe a locally elevated

segment of the Moho surrounded by deeper segments. This appears as an imbrication

in the 2-D profile (Fig. 6), while its 3-D shape is an elongated dome (Fig. 8). Acton

et al. [2011] reported a southward dipping Moho based on a 2-D study, consistent with

our observed geometry. Beyond GTOK, however, the Moho dips northward with minor

lateral variations. Beneath MANG, the Moho is at a depth of ∼54 km. Beneath CHTG

it varies from 52–56 km, progressively dipping northward. Beneath LCHG, we observe

significant lateral variations with a deeper Moho (∼60 km) in the south and the east, while

shallower segments (∼56 km) are observed to the north and the west. As a result, we

observe southward and eastward dipping segments of the Moho beneath LCHG (similar

to Fig. 5a). In 2-D profile (Fig. 6), the eastward dip is masked due to averaging, but

the southward dip is observable and could be the result of imbrication of the lowermost

crust. The average Moho depths beneath GTOK, CHTG and LCHG obtained in this

study matches with the results of Acton et al. [2011]. In addition, our study highlights

the lateral variation in the crustal structure.

The mid-crustal layer is also observed to vary laterally beneath the Higher Himalaya.

Beneath GTOK, it is at a depth of ∼27 km) in the northern back-azimuth and deepens to

∼36 km to the south, indicating a southward dip (Fig. 6 and Fig. 5b). For MANG, the

mid-crustal layer dips gently northward from a depth of ∼29 km to ∼43 km. Further north,

beneath CHTG, the mid-crustal layer shallows to ∼30 km in the southern back-azimuth,

and then continues to dip northward reaching a depth of ∼37 km. This results in a local

2017
c American Geophysical Union. All Rights Reserved.
increase of the northward dipping angle to ∼16◦ from ∼10◦ . Beneath LCHG, a shallow

segment (∼37 km) is seen to the east of the station. In all other directions, the mid-crustal

layer is at a depth of ∼41 km. The average VS of the upper crust increases gradually from

∼3.45 km/s to ∼3.63 km/s, while the average VS of the lower crust reduces from ∼3.8 km/s

to ∼3.7 km/s northwards. This increase in VS of the upper crust northward indicates

mafic composition. The reduction in velocity of the lower crust could be influenced by

the presence of oblique transverse structures beneath the Higher Himalaya, as observed

in the CCP stack (Fig. 2c and 6). The average VS of the entire crust is modeled to

be ∼3.63 km/s and that of the uppermost mantle is ∼4.22 km/s, similar to the Lesser

Himalaya.

3.3.3. Tethyan Himalaya

Although YUMT is located in the Higher Himalaya, the receiver function ray-paths

converging at the station mostly sample the crust beneath the Tethyan Himalaya. Our

joint inversion results indicate that the MHT dips steeply beneath this region. Its depth

increases from ∼19 km in the south to ∼25 km in the north with a dip of ∼16◦ . The

average VS of the MHT LVL is modeled to be ∼3.31 km/s. This is slightly higher than

that in the Lesser and Higher Himalaya, but is nearly equal to the VS in the Himalayan

wedge (∼3.33 km/s). This gradual increase in the LVL VS from the Lesser Himalaya to the

Tethyan Himalaya is probably a result of the transition from frictionally locked to creeping

behavior of the MHT decollement. An additional low-velocity layer is also observed just

above the MHT. This corresponds to the MCT, which splays from the MHT further to the

north. The Moho is at a depth of ∼59 km beneath YUMT, with a marginally elevated

segment (∼56 km) in the NE back-azimuth (Fig. 8). The mid-crustal layer could not

2017
c American Geophysical Union. All Rights Reserved.
be identified by joint-inversion modelling in two clusters of YUMT. This may be due to

destructive interference of the mid-crustal phase with multiples from shallow layers in the

Himalayan wedge. This hypothesis has also been confirmed by forward modelling. In the

remaining clusters, the depth to the mid-crustal layer drastically increases from ∼35 km to

∼47 km northwards, resulting in a tapered lower crust, also observed in previous studies

[Mitra et al., 2005]. The average VS of the lower crust beneath the Tethyan Himalaya

is ∼3.76 km/s, which is higher than the Higher Himalaya. However, the upper crust

beneath both regions has similar average VS (∼3.63 km/s). As a result, the average VS

of the entire crust beneath the Tethyan Himalaya increases slightly to ∼3.67 km/s. The

average VS of the uppermost mantle is modeled to be ∼4.26 km/s, similar to the Lesser

and Higher Himalaya.

To first order, the MHT, mid-crustal interface and Moho dip to the north at angles of

∼9◦ ∼10◦ and ∼7◦ , respectively. However, there are observable second order variations

in the depth of each of these discontinuities. The MHT shows a ramp-flat-ramp geometry

with a gentle northward dip beneath the Lesser Himalaya and a steep dip beneath the

Tethyan Himalaya. Imbricate faults (MCT and PT) splay off the MHT beneath the

Higher Himalaya. The Moho has several secondary undulations with significant southward

dips beneath the Lesser and the Higher Himalaya and lateral heterogeneity beneath the

Lesser Himalaya. The mid-crustal layer also displays southward dips and lateral variations

similar to the Moho. The thickness of the Indian crust (between MHT and the Moho)

varies between ∼35 km and ∼42 km, which is similar to the thickness variation observed

in the Indian Shield crust. Beneath the Tethyan Himalaya the underthrust Indian lower

crust is seen to be marginally tapered. A number of oblique low-velocity layers, within

2017
c American Geophysical Union. All Rights Reserved.
the underthrust Indian crust, are observed beneath the Higher Himalaya (Fig. 6). These

layers are near-perpendicular to the profile and therefore, are oriented NW-SE. They dip

at an high angle to the first order crustal discontinuities and extend from the MHT to the

Moho, intersecting the entire underthrust Indian crust. These structures are discussed in

detail later.

3.4. 3-D mapping of sub-surface structures

The results of the CCP stack, H-K stack and joint inversion highlight the presence

of significant lateral variation in the geometry of the MHT and the Moho beneath the

Sikkim Himalaya. We use the shear-wave velocity model obtained from joint inversion to

map the depths of these discontinuities. This allows us to image the 3-D variation in the

subsurface structures.

3.4.1. Geometry of The Main Himalayan Thrust

Following the procedure outlined in Section 2.2, we map the 3-D configuration of the

MHT. Our results are plotted in Fig. 7, viewed from different directions, with the 3-D

MHT surface being colour-coded by depth. The piercing points of receiver functions are

plotted as black asterisks and stations are plotted as white triangles. The piercing points

from four stations NAMC, RABN, GTOK and MANG constrain the surface beneath the

Lesser Himalaya (Fig. 7a). Each of these stations have four clusters of piercing points.

On closer inspection it is evident that each of the stations has at least two sets of piercing

points at different depths which indicates that the receiver functions sample the edges or

slopes of the 3-D surface of the discontinuity. To highlight the structure defined by the

surface, we draw dashed isolines of depth on the map view (Fig. 7a) and plot a schematic

(Fig. 7b) based on piercing points. The shallow (8–9 km) isolines are drawn in white and

2017
c American Geophysical Union. All Rights Reserved.
the deeper (10–12 km) ones in blue. These are connected by cyan lines. We observe two

oblique ramps (Fig. 7a, b), one dipping to the NE and the other dipping S-SE. However,

we cannot rule out the possibility of these oblique ramps being a combination of number

of ramps with different orientations. The same is true for the flat top portion which

could be rounded or of any irregular shape. Each of these combinations, however, would

result in a dome shape of the MHT beneath the Lesser Himalaya. To investigate the 3-D

structure of the MHT we view the block from two different angles. Viewing the block from

the E-SE direction (Fig. 7c) highlights the ramp-flat-ramp geometry of the MHT, with

the shallower ramp beneath the Lesser Himalaya and the deeper ramp underneath the

Tethyan Himalaya. The two ramps are connected by a flat (dip ∼5◦ ). The ramp beneath

the Lesser Himalaya is smaller and has a gentler dip of ∼7◦ , while the one beneath the

Tethyan Himalaya is larger and has a steeper dip of ∼15◦ . Viewing the block from the

NE (Fig. 7d) shows that the ramp beneath the Lesser Himalaya is domed in shape. This

perspective also highlights the steep dip of the ramp beneath the Tethyan Himalaya. The

arcuate shape of the PT and MCT on the surface (Fig. 7d) matches the spatial extent of

this dome structure on the MHT in the subsurface. We conjecture that the arcuate shape

of the PT and the MCT (observed on regional outcrop map) is the surface expression of

this sub-surface dome-shaped geometry of the MHT. The PT and the MCT splay off the

MHT close to where the two ramps have been observed (Fig. 7d). From this geometry,

we infer that these splays will also be arcuate in shape in the subsurface. Our inference

is supported by the observed dips of the PT and the MCT measured on outcrops, which

reveal that the structure is a doubly plunging anticline (dome shape) [Bhattacharyya and

Mitra, 2009]. Moreover, Ray [2000] demonstrated that the observed ∼60 km N-S spatial

2017
c American Geophysical Union. All Rights Reserved.
deflection of the arcuate shape of the PT and MCT cannot be produced by cross-cutting

of topography and planar thrust sheets, which is consistent with our results.

3.4.2. Geometry of the Moho

The 3-D geometry of the Moho (Fig. 8) is obtained in a similar way to that of the MHT.

The precision and reliability of the 3-D surface constructed from the piercing points are

tested in the same way. The CCP stack results (Fig. 6) suggest that the Moho dips

to the north from the Lesser Himalaya. However, the 3-D surface map shows that the

significant dip of the Moho initiates beneath the Higher Himalaya and continues beyond

the Tethyan Himalaya. Beneath the Lesser Himalaya and the Higher Himalaya, we observe

a significant dome structure on the Moho. This dome structure is slightly staggered from

the dome in the MHT but within the arcuate shape of the PT and the MCT. The crest

of the Moho dome lies below the stations GTOK and MANG. Similar but smaller domes

are also observed in south-eastern (88.9◦ E, 27.3◦ N) and south-western (88.2◦ E, 27.3◦ N)

edges of the map. The CCP stack along an oblique profile overlapped these dome shapes

in the Lesser and the Higher Himalaya to form an apparent double-Moho (Fig. 6). In the

NE corner of the map, we observe that the Moho shallows to a depth of 55–56 km. This

shallow portion is surrounded by deeper segments of the Moho. This could be the remnant

signature of a dome which is now steeply dipping beneath the Tethyan Himalaya. On a

larger scale, these domes are intersected by linear, deeper segments of the Moho trending

NW-SE. The significance of these alternating shallow and deep segments of the Moho is

discussed along with the transverse structures in the lower crust in Section 4.2.

4. Discussion

2017
c American Geophysical Union. All Rights Reserved.
4.1. Crustal architecture of the MHT and its relationship to the surface

structures
We present a detailed image of the Himalayan wedge, the underthrusting Indian crust

and the detachment surface (MHT) along which the Indian plate underthrusts the Sikkim

Himalaya. The top of the underthrusting plate is highlighted by a LVL which marks the

Main Himalayan Thrust [Acton et al., 2011; Nábělek et al., 2009; Schulte-Pelkum et al.,

2005]. This represents a fault zone which facilitates the relative motion between the

Indian plate and the overriding Himalayan wedge. It is observed that the Himalayan

wedge increase in thickness northwards and has an average VS of ∼3.33 km/s. From

our models, we observe that the MHT LVL has a thickness of ∼2 km and an associated

velocity of 3.1–3.2 km/s for stations south of the MCT zone. North of the MCT zone, this

velocity increases and reaches ∼3.31 km/s beneath the Tethyan Himalaya. This translates

into a velocity reduction of ∼15% and ∼10% of the average shear-wave velocity of the

underthrust Indian crust, south and north of the MCT zone, respectively. This change

in the velocity of the MHT LVL may be indicative of the change in rheology of the MHT

from the region of frictionally locked brittle failure to the aseismically creeping zone.

The presence of metamorphic fluids [Patro and Harinarayana, 2009] and/or sediments

under shear stress in the fault zone and/or reduction in grain size due to repeated brittle

failure associated with megathrust earthquakes, over millions of years, can greatly reduce

VS within the locked zone. Additionally, brittle failure in fault zone results in lateral

heterogeneities which act as scatterers for extrinsic attenuation of the shear-wave energy

[Thirunavukarasu et al., 2017]. In contrast, in the creeping zone, VS is reduced only

due to plastic behavior of the materials within the ductile shear zone. The absence of

fluids and extrinsic scatterers leads to lesser reduction of VS in creeping zones compared

2017
c American Geophysical Union. All Rights Reserved.
to the locked zone. Our results suggest that the transition to the creeping zone initiates

somewhere underneath LCHG and YUMT. This is in agreement with the observed locking

line beneath the Sikkim Himalaya [Ader et al., 2012] and coincides with the position of

the deeper ramp on the MHT (Fig. 7), as observed in the Nepal Himalaya [Elliott et al.,

2016].

To first order, the MHT dips gently to the north at ∼9◦ . It is first observed at a depth

of ∼6–8 km beneath the Lesser Himalaya and deepens to ∼25 km beneath the Tethyan

Himalaya. There are observable changes in dip on the MHT, the first one beneath the

Lesser Himalaya and the second one beneath the Tethyan Himalaya. This represents

a ramp–flat–ramp geometry of the MHT, which had also been observed in other parts

of the Himalaya [Mitra et al., 2005; Rai et al., 2006; Caldwell et al., 2013]. The ramp

beneath YUMT is steeper (∼16◦ ) and is connected to the Lesser Himalayan ramp by a

flat (dip ∼5◦ ). A few LVLs are observed above the MHT within the Lesser and the Higher

Himalaya. We conjecture these to be the PT and the MCT which splayed off the MHT.

From our 3-D image of the MHT, we observe a dome structure formed by a combination

of ramps beneath the arcuate MCT culmination zone (Fig. 7). The dips of the ramps are

gentle (∼8◦ ). The presence of this dome upwarps the overlying structure. The PT and

MCT, which appear to originate from the ramps north of this dome structure, are also

upwarped.

The doming up of the MHT could result from elevated topography on the underthrusting

Indian plate. In this case, the thickness of the underthrust Indian crust beneath the

dome structure should be higher than the surrounding crust, and this is observed from

our inverted models. In view of this, we consider the MHT dome to be analogous to a

2017
c American Geophysical Union. All Rights Reserved.
subducting seamount on the top of a downgoing oceanic plate in a subduction system.

Upwarping of overlying strata and heterogeneous and complex faulting is observed in the

vicinity of seamounts [Dominguez et al., 1998]. By a conceptual model, Wang and Bilek

[2011] showed that the horizontal tensile stress exerted by a seamount on the upper plate

is one order of magnitude greater than the normal stress it exerts. Therefore, instead

of producing more friction to create large earthquakes, the upper plate crumbles to a

network of numerous fractures. This system of failed faults acts as a creep zone, allowing

the seamount to propagate with very little or no hindrance resulting in aseismic creeping,

but can also be associated with microseismicity. As the Indian plate underthrusts, this

dome structure in the lesser Himalaya creates a network of fractures which may have

initiated the active Lesser Himalayan Duplex. In the case of the subducting seamount,

microseismicity is confined to a region adjacent to the seamount while the smoother regions

are frictionally locked. Similar observations have been made in the regions adjacent to the

Lesser Himalayan dome structure. From microseismicity studies in the Sikkim Himalaya

[De and Kayal , 2003; Hazarika et al., 2010], we observe that the seismicity is clustered

around the region of this dome structure. As we go away from the dome, the frequency of

microseismicity decreases. The dome structure may also act like a seamount in stopping

rupture propagation when the smooth patch of the neighboring locked zone breaks [Wang

and Bilek , 2011]. Evidence of such rupture termination of megathrust earthquakes has

been seen in the Java subduction zone [Bilek and Engdahl , 2007]. The rupture area of the

1934 Bihar-Nepal earthquake (Mw 8.1) did not propagate eastward of the Nepal–Sikkim

border [Sapkota et al., 2013]. The presence of this dome structure underneath Sikkim may

provide an explanation for this eastward rupture termination. Fault zones, where some

2017
c American Geophysical Union. All Rights Reserved.
patches rupture and some parts inhibit rupture, are known to produce low stress drops

[Brown et al., 2015; Wang and Bilek , 2011]. Faulting of the strong segment produces

large stress drop but simultaneously increases the stress in the neighboring regions. As a

result, the average stress drop is observed to be low. In Sikkim, the stress drop associated

with moderate-to-large earthquakes were found to be low [Paul et al., 2015] supporting

the notion of heterogeneous faulting.

4.2. Architecture of the underthrust Indian crust and its relationship to

transverse tectonics
The mid-crustal discontinuity, within the underthrust Indian plate, is consistently ob-

served beneath the Sikkim Himalaya. This interface has a first order dip to the north and

also displays lateral heterogeneity in the form of steep dips (beneath RABN and CHTG),

southward dip (beneath GTOK) and imbrication (beneath CHTG). The thickness of the

lower Indian crust (between the mid-crustal layer and the Moho) lies between ∼14–25 km.

The average VS of the lower crust decreases from ∼3.8 km/s to ∼3.7 km/s as we go north-

wards from the Lesser Himalaya to the Higher Himalaya. A similar observation has been

made in the Uttaranchal Himalaya by Caldwell et al. [2009], which they ascribed to the

presence of aqueous fluids/partial melts in the lower crust. We observe oblique LVLs

beneath the Higher Himalaya and identify them as transverse structures (faults) [Paul

et al., 2015]. These low velocity transverse structures may provide an alternative expla-

nation for the reduction in velocity. Such structures can also facilitate the propagation of

fluids/melts from the lower crust. The average VS of the underthrust Indian plate lies be-

tween ∼3.60 km/s and ∼3.65 km/s and its average VP /VS is <1.74, consistent with a felsic

composition. However, the average VS of the lower crust suggests that it is slightly more

2017
c American Geophysical Union. All Rights Reserved.
mafic and is possibly composed of dry granulites. The uppermost mantle (Sn ) velocity

lies between ∼4.15 and ∼4.27 km/s, which is slightly lower than the observed Sn velocity

beneath the Indian Shield [Mitra et al., 2006]. The thickness of the underthrust Indian

plate (between MHT and Moho) varies between 35 km and 42 km, which is consistent

with the average thickness of the Indian Shield crust south of the Himalaya.

Similar to the MHT, the Moho also dips gently to the north at an angle of ∼7◦ . However,

there are a number of second order lateral variations on the Moho (Fig. 8) in the form

of dome structures and deeper sections. These are explained either by imbrication of the

lower boundary of an elastic layer under flexure (Fig. 6) or by transverse faulting through

the lower Indian crust. Paul et al. [2015] studied the source mechanisms of lower crustal

earthquakes beneath the Sikkim Himalaya and demonstrated that the underthrust Indian

crust deforms by strike-slip motion on a set of transverse structures. Such transverse

structures in the lower crust of the Indian plate are also observed in our CCP stacks

(Fig. 6). These structures trend NW–SE and intersect the entire lower crust at regular

intervals (Fig. 8). The 2011 Sikkim earthquake, its deeper aftershocks and previous

strike-slip earthquakes (Fig. 8), in southern Tibet immediately north of our study region,

occurred close to the Moho and ruptured the base of the seismogenic Indian lower crust.

The dextral strike-slip motion associated with this faulting displaces the eastern block of

the lower crust southward relative to the western block. This results in the juxtaposition of

a deeper segment of the Moho to the east with a shallower segment of the Moho to the west

across the fault. The observed undulations on the Moho underneath the Lesser Himalaya

reflect Moho offsets accumulated over several such dextral strike-slip earthquakes (similar

to the Mw 6.9 2011 Sikkim earthquake) on a set of transverse structures present within

2017
c American Geophysical Union. All Rights Reserved.
the underthrust Indian crust beneath the Sikkim Himalaya. The source mechanisms

of historical earthquakes, the 2011 Sikkim mainshock and its deeper aftershocks (Fig. 8)

reveal that these occurred near the edges of these dome structures. These active transverse

structures, along with the distribution of earthquakes, highlight that the lower Indian crust

beneath the Sikkim Himalaya is seismogenic [Mitra et al., 2005; Acton et al., 2011]. The

presence of these active transverse structures should be included in kinematic models of

active deformation of the Eastern Himalayan plate boundary system.

A comparison of the results from the Sikkim Himalaya has been made with the adja-

cent Nepal Himalaya [Schulte-Pelkum et al., 2005; Nábělek et al., 2009] and the Bhutan

Himalaya [Singer et al., 2017]. The geometry of the Himalayan detachment (MHT) be-

neath the eastern Nepal Himalaya [Schulte-Pelkum et al., 2005] has a gentle northward

dip, unlike the ramp-flat-ramp geometry observed beneath Sikkim. Further west beneath

the central Nepal Himalaya [Nábělek et al., 2009], the gentle north dipping MHT has been

observed to display changes in dip at two points, the first below the Lesser Himalaya and

the second below the Higher-to-Tethyan Himalaya. These are similar to the 2-D variation

observed on the MHT beneath the Sikkim Himalaya. The thickness of the underthrust

Indian crust beneath the Nepal Himalaya is comparable to the Sikkim Himalaya, but the

geometry of the Moho is relatively smooth and dips to the N-NE. Beneath the Bhutan

Himalaya (east of Sikkim) [Singer et al., 2017], the MHT and Moho have lateral variation

from the west to the east, similar to Sikkim. A significant difference in results lies beneath

the western Bhutan, where the Moho reveals a steep dip south of the Higher Himalaya

and reaches a depth of ∼70 km. The MHT too has a moderate change in dip beneath the

2017
c American Geophysical Union. All Rights Reserved.
Lesser-to-Higher Himalaya and appears as an eastward continuation of the ramp observed

beneath the Sikkim Himalaya.

5. Conclusions

We study the 3-D architecture of the crust beneath the Sikkim Himalaya using receiver

function analysis and joint inversion of receiver function and Rayleigh-wave group ve-

locity dispersion data. The Indian plate underthrusts the Sikkim Himalaya with its top

(MHT) highlighted by a LVL and its base (Moho) by a large impedance contrast bound-

ary observed in receiver functions from all the stations. To first order all the crustal

discontinuities – the MHT, the mid-crust and the Moho dip northward, at ∼9◦ , ∼10◦ and

∼7◦ , respectively. From detailed images, we observe that the MHT has a ramp-flat-ramp

geometry beneath the Lesser and Tethyan Himalaya in Sikkim. The Lesser Himalayan

ramp forms a dome structure on the MHT. This upwarps the overlying PT and MCT

to form an arcuate geometry observed in its outcrop. This dome geometry on the MHT

is a result of elevated topography on the underthrust Indian plate and is analogous to

subducting sea-mounts. Such a dome structure has the potential to terminate rupture

propagation of megathrust earthquakes and result in low stress drops. MHT LVL has a

reduction of VS by 15% and 10% beneath the Lesser Himalaya and the Tethyan Himalaya,

respectively. This northward increase in VS within the LVL possibly marks the transition

from frictionally locked brittle regime to freely sliding aseismic creep regime.

The average VS of the entire Indian crust is ∼3.63 km/s and its VP /VS is ∼1.72±0.02

which indicates a felsic crustal composition. The average VS of the upper crust increases

marginally from south to north, while that of the lower crust decreases northward. We

conjecture that this decrease in VS of the lower crust is either due to the variation in the

2017
c American Geophysical Union. All Rights Reserved.
crustal structure of the underthrust Indian plate or due to the presence of active transverse

structures beneath the Higher Himalaya. The average crustal thickness of the underthrust

Indian crust varies between 35 km and 42 km and is consistent with the average thickness

of the Indian Shield crust south of the Himalaya. The mid-crustal discontinuity displays

lateral heterogeneity in the form of steep, southward dip and imbrications. The Moho

has second order lateral variations in the form of dome structures and undulations. These

result from imbrication of the lower boundary of an elastic layer under flexure and/or

active transverse faulting in the lower crust.

Acknowledgments

We thank the Department of Science and Technology, India (DST) Deep Continental

Studies (DCS) program for funding the equipment used in the temporary broadband net-

work in Sikkim and UK-India Education and Research Initiative (UK-IERI) Thematic

Partnership 2011 for providing the operational cost of the experiment between 2011-

2014. SM thanks Keith Priestley (University of Cambridge) for sharing the previously

collected data (2005) from the GTOK station. All receiver function data used in this

study are available through the IISER Kolkata Seismological Observatory data repository

at the following url http://www.iiserkol.ac.in/∼supriyomitra/Sikkim RFs.tgz . Support

from Varun Joshi during the field experiment is acknowledged. We also acknowledge G. B.

Pant Institute of Himalayan Environment and Development (GBPIHED), Pangthang and

their personnel for lending support in operating the GTOK station. Data pre-processing

and part of the analysis was performed using Seismic Analysis Code, version 100 [Gold-

stein et al., 2003]. All plots were made using the Generic Mapping Tools version 4.0

(www.soest.hawaii.edu/gmt); Wessel and Smith [1998]. HP acknowledges PhD scholar-

2017
c American Geophysical Union. All Rights Reserved.
ship from IISER Kolkata. SM thanks research funding from DCS-DST, India (2008-2011),

UK-IERI Thematic Partnership (2011) and Academic Research Funding (ARF), IISER

Kolkata. We thank Camilla Penny for critically reading the manuscript and two anony-

mous referees and the associate editor for their critical reviews and constructive comments

which has helped us to improve the manuscript significantly.

References

Acton, C., K. Priestley, S. Mitra, and V. Gaur (2011), Crustal structure of the Darjeeling–

Sikkim Himalaya and southern Tibet, Geophys. J. Int., 184, 829–852.

Ader, T., J.-P. Avouac, J. Liu-Zeng, H. Lyon-Caen, L. Bollinger, J. Galetzka, J. Genrich,

M. Thomas, K. Chanard, S. N. Sapkota, et al. (2012), Convergence rate across the Nepal

Himalaya and interseismic coupling on the Main Himalayan Thrust: Implications for

seismic hazard, J. Geophys. Res, 117 (B4).

Avouac, J.-P. (2003), Mountain building, erosion, and the seismic cycle in the Nepal

Himalaya, in Advances in Geophysics, vol. 46, edited by R. Dmowska, pp. 1–88.

Avouac, J.-P., and P. Tapponnier (1993), Kinematic model of active deformation in Cen-

tral Asia, Geophy. Res. Lett., 20 (10), 895–898.

Bhattacharyya, K., and G. Mitra (2009), A new kinematic evolutionary model for the

growth of a duplex – an example from the Rangit duplex, Sikkim Himalaya, India,

Gondwana Res., 16 (3), 697.

Bilek, S., and E. Engdahl (2007), Rupture characterization and aftershock relocations for

the 1994 and 2006 Tsunami earthquakes in the Java subduction zone, Geophy. Res.

Lett., 34, L20,311.

2017
c American Geophysical Union. All Rights Reserved.
Brown, L., K. Wang, and T. Sun (2015), Static stress drop in the Mw 9 Tohoku-oki

earthquake: Heterogeneous distribution and low average value, Geophy. Res. Lett., 42,

10595–10600.

Caldwell, W., S. Klemperer, S. S. Rai, and J. F. Lawrence (2009), Partial melt in the

upper-middle crust of the northwest Himalaya revealed by Rayleigh wave dispersion,

Tectonophys., 477 (1–2), 58–65.

Caldwell, W., S. Klemperer, J. F. Lawrence, S. S. Rai, and Ashish (2013), Characterizing

the Main Himalayan Thrust in the Garhwal Himalaya, India with receiver function CCP

stacking, Earth Planet. Sci. Lett., 367, 15–27.

De, R., and J. Kayal (2003), Seismotectonic model of the Sikkim Himalaya: Constraint

from microearthquake surveys, Bull. Seismol. Soc. Am., 93 (3), 1395–1400.

Dominguez, S., S. Lallem, J. Malavielle, and R. von Huene (1998), Upper plate deforma-

tion associated with seamount subduction, Tectonophys., 293, 207–224.

Elliott, J., R. Jolivet, P. Gonzalez, J.-P. Avouac, J. Hollingsworth, M. Searle, and

V. Stevens (2016), Himalayan megathrust geometry and relation to topography revealed

by the Gorkha earthquake, Nat. Geosci., 9, 174–180.

England, P., and G. Houseman (1986), Finite strain calculations of continental deforma-

tion: 2. comparison with the India-Asia Collision Zone, J. Geophys. Res., 91.

England, P., and P. Molnar (1997), Active Deformation of Asia: From Kinematics to

Dynamics, Science, 278, 647–650.

Gilligan, A. (2014), Imaging the structure of the crust and upper mantle in central Asia,

Ph.D Thesis, University of Cambridge, Cambridge, UK.

2017
c American Geophysical Union. All Rights Reserved.
Goldstein, P., D. Dodge, M. Firpo, and L. Minner (2003), Sac2000: Signal processing and

analysis tools for seismologists and engineers., in The IASPEI International Handbook

of Earthquake and Engineering Seismology, edited by W. Lee, H. Kanamori, P. C.

Jennings, and C. Kisslinger, Academic Press, London.

Hazarika, P., M. Ravi Kumar, G. Srijayanthi, P. Solomon Raju, N. Purnachandra Rao,

and D. Srinagesh (2010), Transverse tectonics in the Sikkim Himalaya: Evidence from

seismicity and focal-mechanism data, Bull. Seismol. Soc. Am., 100, 1816–1822.

Herrmann, R., and C. Ammon (2004), Surface waves, receiver functions and crustal struc-

ture. Computer Programs in Seismology, Version 3.30 [electronic]. Saint Louis Univ, St.

Louis, MO.

Hirn, A., J. Lepine, G. Jobert, et al. (1984), Crustal structure and variability of the

Himalayan border of Tibet, Nature, 307 (5946), 23–25.

Kumar, A., S. Mitra, and G. Suresh (2015), Seismotectonics of the Eastern Himalayan

and Indo-Burman plate boundary systems, Tectonics, 34 (11), 2279–2295.

Lei, J., and D. Zhao (2016), Teleseismic P-wave tomography and mantle dynamics beneath

Eastern Tibet, Geochem. Geophys. Geosyst, 17, 1861–1884.

Lei, J., F. Xie, J. Teng, G. Zhang, C. Sun, and X. Zha (2014), Pn anisotropic tomography

and dynamics under eastern Tibetan plateau, J. Geophys. Res., 119, 2174–2198.

Li, C. R., v.-d. Hilst, A. S. Meltzer, and E. R. Engdahl (2008), Subduction of the Indian

lithosphere beneath the Tibetan Plateau and Burma, Earth Planet. Sci. Lett., 274,

157–168.

Ligorria, J., and C. Ammon (1999), Iterative deconvolution and receiver-function estima-

tion, Bull. Seismol. Soc. Am., 89, 1395–1400.

2017
c American Geophysical Union. All Rights Reserved.
Mitra, S., K. Priestley, A. K. Bhattacharyya, and V. K. Gaur (2005), Crustal structure

and earthquake focal depths beneath northeastern India and Southern Tibet, Geophys.

J. Int., 160, 227–248.

Mitra, S., K. Priestley, V. K. Gaur and S. S. Rai (2006), Shear-Wave Structure of the South

Indian Lithosphere from Rayleigh Wave Phase Velocity Measurements, Bull. Seismol.

Soc. Am., 96 (4), 1551–1559.

Nábělek, J., G. Hetényi, J. Vergne, S. Sapkota, B. Kafle, M. Jiang, H. Su, J. Chen, B.-S.

Huang, et al. (2009), Underplating in the Himalaya-Tibet collision zone revealed by the

Hi-CLIMB experiment, Science, 325 (5946), 1371–1374.

Ni, J. F., M. Guzman-Speziale, M. Bevis, W. E. Holt, T. C. Wallace, and W. R. Seager

(1989), Accretionary tectonics of Burma and the three-dimensional geometry of the

Burma subduction zone, Geology, 17 (1), 68–71.

Patro, P. K., and T. Harinarayana (2009), Deep geoelectric structure of the Sikkim Hi-

malayas (NE India) using magnetotelluric studies, Phys. Earth Planet. Inter., 173 (1),

171–176.

Paul, H., S. Mitra, S. Bhattacharya, and G. Suresh (2015), Active transverse faulting

within underthrust Indian crust beneath the Sikkim Himalaya, Geophys. J. Int., 201 (2),

1070–1081.

Rai, S., K. Priestley, V. Gaur, S. Mitra, M. Singh, and M. Searle (2006), Configuration of

the Indian moho beneath the NW Himalaya and Ladakh, Geophy. Res. Lett., 33 (15).

Ray, S. K. (2000), Culmination zones in Eastern Himalaya, in Proceedings. Dr. M. S.

Krishnan Birth Centenary Seminar, vol. 55, pp. 85–94, Geol. Survey of India.

2017
c American Geophysical Union. All Rights Reserved.
Sapkota, S. N., L. Bollinger, Y. Klinger, P. Tapponnier, Y. Gaudemer, and D. Tiwari

(2013), Primary surface ruptures of the great Himalayan earthquakes in 1934 and 1255,

Nature, 6, 71–76.

Schulte-Pelkum, V., G. Monsalve, A. Sheehan, M. Pandey, S. Sapkota, R. Bilham,

and F. Wu (2005), Imaging the Indian subcontinent beneath the Himalaya, Nature,

435 (7046), 1222–1225.

Singer, J., E. Kissling, T. Diehl, and G. Hetényi (2017), The underthrusting Indian crust

and its role in collision dynamics of the Eastern Himalaya in Bhutan: Insights from

receiver function imaging, J. Geophys. Res., 122 (2), 1152–1178.

Singh, A., M. R. Kumar, and P. S. Raju (2010), Seismic structure of the underthrusting

Indian crust in Sikkim Himalaya, Tectonics, 29 (6).

Thirunavukarasu, A., A. Kumar, and S. Mitra (2017), Lateral variation of seismic atten-

uation in Sikkim Himalaya, Geophys. J. Int., 208 (1), 257–268.

Tiwari, V., M. V. Rao, D. Mishra, and B. Singh (2006), Crustal structure across Sikkim,

NE Himalaya from new gravity and magnetic data, Earth Planet. Sci. Lett., 247 (1),

61–69.

Wang, K., and S. Bilek (2011), Do subducting seamounts generate or stop large earth-

quakes?, Geology, 39 (9), 819–822.

Wessel, P., and W. H. F. Smith (1998), New, improved version of the generic mapping

tools released., EOS Trans. AGU, 79, 579.

Yin, A. (2000), Mode of Cenozoic east-west extension in Tibet suggesting a common origin

of rifts in Asia during the Indo-Asian collision, J. Geophys. Res., 105, 21,745–21,759.

2017
c American Geophysical Union. All Rights Reserved.
Yin, A. (2006), Cenozoic tectonic evolution of the Himalayan orogen as constrained by

along-strike variation of structural geometry, exhumation history, and foreland sedimen-

tation, Earth-Sci. Reviews, 76, 1–131.

Yin, A., and T. M. Harrison (2000), Geologic evolution of the Himalayan Tibetan orogen,

Annu. Rev. Earth Planet. Sci., 28, 211–280.

Zhu, L. (2000), Crustal structure across the San Andreas Fault, southern California from

teleseismic converted waves, Earth Planet. Sci. Lett., 179, 183–190.

Zhu, L., and H. Kanamori (2000), Moho depth variation in southern California from

teleseismic receiver functions, J. Geophys. Res., 105, 2969–2980.

2017
c American Geophysical Union. All Rights Reserved.
Table 1. Summary of receiver function back-azimuth clusters and the H-K stacking results.
No. Station Clust. BAZ No. of Mid-crust Moho
No. (◦ ) RFs
Depth VP /VS Depth VP /VS
(km) (km)
1. NAMC 1 30-45 9 37.5±2.66 1.690±0.030 48.8±2.66 1.765±0.030
2 54-72 13 36.0±1.47 1.720±0.010 50.0±1.47 1.690±0.010
3 80-95 7 36.2±3.66 1.718±0.031 – –
4 105-157 26 32.5±2.18 1.675±0.015 52.0±2.18 1.736±0.015
5 295-313 6 34.5±4.18 1.750±0.041 43.0±4.18 1.785±0.041
2. RABN 1 33-54 36 33.5±2.03 1.700±0.014 46.1±2.03 1.705±0.014
2 55-59 17 27.0±4.40 1.760±0.031 49.0±4.40 1.640±0.031
3 60-72 74 32.5±1.32 1.690±0.010 48.5±1.32 1.660±0.010
4 74-91 19 30.5±1.38 1.720±0.016 44.0±1.38 1.725±0.016
5 94-105 14 26.5±2.16 1.720±0.018 45.0±2.16 1.680±0.018
6 106-116 75 27.5±1.29 1.690±0.006 45.5±1.29 1.690±0.006
7 117-126 29 27.5±1.69 1.695±0.014 46.5±1.69 1.690±0.014
8 127-157 32 27.5±1.84 1.700±0.014 43.0±1.84 1.665±0.014
9 277-308 24 30.0±2.86 1.660±0.013 50.0±2.86 1.700±0.013
3. GTOK 1 5-47 35 29.0±2.74 1.700±0.016 50.0±2.74 1.775±0.016
2 49-57 21 31.0±4.41 1.700±0.027 55.0±4.41 1.720±0.027
3 59-74 84 30.0±1.17 1.740±0.009 56.5±1.17 1.730±0.009
4 76-98 46 29.0±2.83 1.740±0.013 – –
5 102-152 222 34.0±1.14 1.775±0.006 58.0±1.14 1.710±0.006
6 153-158 86 38.0±1.15 1.700±0.008 49.5±1.15 1.785±0.008
7 235-256 12 29.0±6.52 1.690±0.027 57.5±6.52 1.735±0.027
8 282-305 22 27.0±3.26 1.760±0.024 61.0±3.26 1.750±0.024
4. MANG 1 51-83 7 46.0±8.98 1.700±0.062 61.0±8.98 1.790±0.062
2 107-110 4 44.0±9.86 1.730±0.012 62.5±9.86 1.770±0.012
3 120-149 5 31.5±6.83 1.750±0.040 53.0±6.83 1.675±0.040
4 152-172 6 35.0±6.81 1.690±0.040 55.0±6.81 1.680±0.040
5. CHTG 1 41-96 44 43.0±2.56 1.730±0.011 58.5±2.56 1.710±0.011
2 106-138 71 38.0±1.69 1.710±0.008 58.0±1.69 1.715±0.008
3 147-158 33 33.0±2.49 1.700±0.011 55.0±2.49 1.720±0.011
4 293-306 5 47.0±1.73 1.720±0.017 59.0±1.73 1.710±0.017
6. LCHG 1 29-50 13 44.0±6.04 1.770±0.029 57.0±6.04 1.740±0.029
2 51-97 38 47.0±4.36 1.700±0.014 63.5±4.36 1.720±0.014
3 106-129 45 44.5±4.99 1.700±0.017 59.0±4.99 1.710±0.017
4 134-208 30 44.0±2.57 1.710±0.016 56.5±2.57 1.690±0.016
7. YUMT 1 27-42 18 47.0±1.82 1.690±0.020 58.0±1.82 1.665±0.020
2 45-68 63 – – 53.0±1.87 1.740±0.018
3 70-89 24 – – 59.0±3.17 1.700±0.030
4 105-140 106 42.0±1.75 1.700±0.008 66.2±1.75 1.740±0.008
5 147-158 37 33.0±3.90 1.720±0.012 60.0±3.90 1.750±0.012

2017
c American Geophysical Union. All Rights Reserved.
Table 2. Summary of the joint inversion results. The MHT depths indicated by * are thick

LVL (4–6 km) with the given depth as upper bound of the LVL. These may be a combination

of the MHT and its imbricate faults above. The uncertainty in depth is ±2 km and in VS is

±0.05 km/s.
Station No. BAZ LVL (MHT) Himalayan Indian Indian Indian Upper
(◦ ) Wedge Upper crust Lower Crust Whole Crust Mantle
Depth VS VS Thickness VS Thickness VS Thickness VS VS
of layer of layer of layer of layer of layer of layer of layer of layer of layer Sn
(km) (km/s) (km/s) (km) (km/s) (km) (km/s) (km/s) (km) (km/s)
NAMC 1 30-45 10.5 3.21 3.28 33 3.54 19 3.83 52 3.63 4.23
2 54-72 8 3.15 3.31 33 3.48 12 3.88 45 3.58 4.07
3 80-95 6* 3.16 3.29 27 3.28 24 3.77 51 3.58 4.26
4 105-157 6* 3.13 3.30 27 3.38 24 3.79 51 3.60 4.22
5 295-313 7.5 3.14 3.34 33 3.55 11 3.73 44 3.59 4.19
RABN 1 33-54 7* 3.16 3.26 28 3.41 17 3.90 45 3.62 4.19
2 55-59 6* 3.07 3.37 26 3.35 18 3.88 44 3.58 4.14
3 60-72 7* 3.16 3.26 28 3.41 17 3.90 45 3.62 4.19
4 74-91 6* 3.25 3.31 28 3.42 16 3.88 44 3.60 4.02
5 94-105 10 3.26 3.29 24 3.33 20 3.84 44 3.63 4.33
6 106-116 9.5 3.10 3.22 24 3.39 28 3.78 52 3.65 4.20
7 117-126 9.5 3.06 3.25 24 3.44 29 3.83 53 3.69 4.26
8 127-157 9.5 3.10 3.25 24 3.48 27 3.84 51 3.69 4.07
9 277-308 10.5 3.04 3.31 25 3.38 27 3.83 52 3.69 4.26
GTOK 1 5-47 10 3.15 3.24 25 3.32 27 3.76 52 3.60 4.30
2 49-57 10.5 3.19 3.24 28 3.39 26 3.86 54 3.66 4.36
3 59-74 6* 3.29 3.33 28 3.36 26 3.80 54 3.60 4.24
4 76-98 11 3.26 3.30 21 3.25 24 3.66 45 3.57 4.07
5 102-152 12 3.26 3.28 35 3.66 20 3.79 55 3.66 4.23
6 153-158 12 3.28 3.27 36 3.67 18 3.73 54 3.63 4.26
7 235-256 7 3.29 3.30 40 3.54 14 3.85 54 3.63 4.24
8 282-305 8* 3.24 3.28 27 3.35 27 3.85 54 3.63 4.39
MANG 1 51-83 13 3.18 3.32 43 3.63 11 3.93 54 3.67 4.19
2 107-110 13 3.25 3.25 36 3.56 17 3.71 53 3.62 4.21
3 120-149 9 3.25 3.39 29 3.48 25 3.83 54 3.63 4.30
4 152-172 10 3.19 3.25 29 3.44 26 3.79 55 3.64 4.20
CHTG 1 41-96 11* 3.22 3.25 40 3.53 16 3.76 56 3.76 4.24
2 106-138 12.5 3.29 3.25 37 3.59 17 3.67 54 3.57 4.23
3 147-158 15 3.28 3.31 30 3.46 22 3.74 52 3.60 4.22
4 293-306 15 3.13 3.26 36 3.59 19 3.70 55 3.63 4.24
LCHG 1 29-50 16 3.22 3.26 41 3.64 14 3.69 55 3.59 4.21
2 51-97 16 3.33 3.26 42 3.67 16 3.64 58 3.60 4.28
3 106-129 17 3.35 3.26 37 3.57 24 3.72 61 3.65 4.28
4 134-208 16 3.22 3.27 40 3.63 21 3.72 61 3.66 4.24
YUMT 1 27-42 25 3.36 3.34 47 3.56 12 3.96 59 3.72 4.28
2 45-68 25 3.36 3.32 – – – 3.62 54 3.64 4.21
3 70-89 25 3.34 3.35 – – – 3.69 57 3.65 4.17
4 105-140 19 3.36 3.31 37 3.47 24 3.81 61 3.67 4.28
5 147-158 19.5 3.14 3.35 35 3.87 24 3.74 59 3.68 4.37

2017
c American Geophysical Union. All Rights Reserved.
American Geophysical Union. All Rights Reserved.
(a) (b)
TIB
ET STD
INDIA
N MCT
E
P
A
L B
H
U
RT PT T
A
N
MBT
MFT

2017
c
Figure 1. (a) Topographic map of Sikkim highlighting the important Himalayan structures: Main Frontal Thrust (MFT),
Main Boundary Thrust (MBT), Main Central Thrust (MCT), Peling Thrust (PT), Rangit Thrust (RT) and South Tibetan
Detachment (STD). The traces of these Himalayan structures are taken from [Bhattacharyya and Mitra, 2009]. The locations
of the broadband stations within our network in Sikkim are shown as cyan triangle. (b) Map of the globe centered on Sikkim
Himalaya, with location of teleseismic events (red stars) used in this study.
(a) (b)
A’
A’
28˚00'

27˚30'

A
27˚00'

88˚00' 88˚30' 89˚00'

c
(c) 6
Gauss 2.5
4
2

2017
NAMC RABN GTOK MANG CHTG LCHG YUMT
0
A A’

−20

−40

Depth/km
−60

−80

−100
0 20 40 60 80 100 120 A
Distance/km

Figure 2. (a) Map showing the location of stations (coloured triangles) and corresponding receiver function piercing points on the
Moho (plotted as ’+’ with the same colour), computed from the average crustal velocity model obtained from the results of Acton et al.
[2011]. (b) Spatial stack of receiver functions binned by their Moho piercing points at every 1 km interval and plotted along the profile
AA’ (A – 88.25◦ E, 27.05◦ N and A’– 88.90◦ E, 28.0◦ N). The positive amplitudes are coloured black and the negative amplitudes are
coloured grey. The P s times corresponding to the MHT, the mid-crustal interface and the Moho are marked by magenta, green and
yellow dashed lines, respectively. (c) The CCP stack along the profile AA’. The amplitude of the phases are scaled relative to the
amplitude of the direct P p phase, coloured red for positive and blue for negative amplitudes. Horizontal smoothing has been applied
using a Gaussian filter of width 2 km. The MHT, mid-crustal interface and Moho are marked as in (b).

American Geophysical Union. All Rights Reserved.


American Geophysical Union. All Rights Reserved.
RABN Cluster 8 (Baz: 127−157 degree)
(b)
k
Depth (Km)
(a)
25 30 35 40 45 Mid−crust (27.5, 1.70)
RABN
2.50 P27.5s 20030101(001)0
1.75 1.754 7 . 8 7 PpP27.5s
0.075
PpS27.5s
0 7 14 21 28
Time (sec)
0.16
0.16
0.04
0.08
0.12

0.2
0.12

0.08
Vp/Vs

0.08

Mid−crust
0.2

0.04

0.04
0.08

(c)
0.04

1.70 1.70
(27.5, 1.70)
Moho (43.0, 1.67)
0.12

RABN
2.50 20030101(001)0
P43s PpP43s
36.15
Moho 0.075
(43.0, 1.67)
PpS43s
1.65 1.65
0 7 14 21 28
25 30 35 40 45 Time (sec)
Depth (Km)
k
Figure 3. (a) H-K stack contour plot (between H and VP /VS ) for RABN (cluster 8, back-azimuth range 127◦ –157◦ ). The
estimated H-K for the mid-crustal layer and the Moho are marked by a blue and a red cross-wire, respectively. The uncertainty

2017
bounds are shown by shaded rectangles. The H and K values for the mid-crust are 27.5±1.84 and 1.70±0.014 and that of

c
the Moho are 43±1.84 and 1.665±0.014, respectively. (b) Plot of forward modeled receiver function stack corresponding to
the estimated H-K for the mid-crustal layer, and (c) for the Moho. The stack of all receiver functions in the cluster is shown
in blue and that of the calculated synthetics is in red. The converted phase (P s) and the reverberating multiples (P pP s and
P pSs) are labelled on the plots. The station name, Gaussian width, percentage fit and ray parameter are listed on the left
from top to bottom.
(a) (c)
RABN
Cluster 8

MHT

Mid−crust

c
2017
(b)
BAZ 127−157 deg
Moho

Figure 4. Joint inversion result of receiver functions and Rayleigh-wave group velocity dispersion data for RABN (cluster

8, back-azimuth range 127◦ –157◦ ). (a) Plot of fit between the observed dispersion data (black circles with error bars) and the

synthetic dispersion data computed from the minimum interface model (blue) and the minimum layer model (red). (b) Plot of

fit between the observed receiver function stack (black dash) and synthetic receiver function stack computed from the models

plotted in (c). (c) Plot of minimum layer model (blue) and minimum interface model (red) obtained by fitting the data. The

starting mantle half-space model is plotted in grey (dashed line). The inferred MHT, mid-crust and Moho are labeled on the

plot.

American Geophysical Union. All Rights Reserved.


Moho Midcrust
88˚30' 89˚00' 88˚30' 89˚00'
28˚00' 28˚00'
(a) (b)

52
48

56
YUMT YUMT

LCHG LCHG

44
CHTG CHTG

Depth 40

60
27˚30' MANG
52 56 27˚30' MANG
36

32
48
32
GTOK GTOK
RABN RABN

32
NAMC NAMC

32
27˚00' 27˚00'
kms kms
44 48 52 56 60 64 28 32 36 40 44 48
88˚30' 89˚00' 88˚30' 89˚00'
28˚00' 28˚00'
(c) (d)

YUMT YUMT

1.72
1.72
2
1.7

LCHG LCHG

CHTG CHTG
Vp/Vs

27˚30' MANG 27˚30' MANG


1.72
2
1.7
1.72
1.7
2

GTOK GTOK
RABN RABN

2
1.7

NAMC NAMC

2
1.7

27˚00' 27˚00'

1.64 1.68 1.72 1.76 1.80 1.64 1.68 1.72 1.76 1.80
Figure 5. Contour plot of depth to (a) the Moho, and (b) the mid-crustal layer, and associated
VP /VS variation for (c) the whole crust, and (d) the crustal layer above the mid-crustal discontinuity,
beneath the Sikkim Himalaya. The depth and VP /VS for each cluster at a given station is projected on
the map (r,θ) using the central back-azimuth (θ), chosen from the back-azimuth range of the cluster,
and a fixed distance (r, proportional to its depth) from the station. The contoured surface is drawn by
triangulation and the variation in depth and VP /VS is colour-coded. The grey triangles represent the
stations and the crosses represent a group of RF piercing points where actual measurements were made.

2017
c American Geophysical Union. All Rights Reserved.
(a) (d)

A’ −30
S N
28˚00'
NAMC RABN GTOK MANG CHTG LCHG YUMT
−40

−50

27˚30'
−60

Moho depth (km)


Moho from H−K stack
Moho from joint inversion

−70
A Relative distance
27˚00'
Depth <= 20
Depth > 20

88˚00' 88˚30' 89˚00'


−4 −2 0 2 4 Lesser Higher Tethyan

c
(b) 6 (c) 6
4 4
2 2
NAMC RABN GTOK MANG CHTG LCHG YUMT NAMC RABN GTOK MANG CHTG LCHG YUMT

2017
0 A
0A ~3.30km/s PT MCT A’
A’
Him. Wedge ~3.29km/s
MHT

~3.33km/s
−20 −20 ~3.42km/s
~3.45km/s
~3.63km/s

Midcrustal layer

TS ~3.63km/s
−40 −40 ~3.82km/s
~3.8km/s ~3.7km/s

Moho ~3.76km/s

Depth/km

Depth/km
−60 −60
~4.22km/s ~4.22km/s ~4.22km/s
~4.26km/s

UPPER MANTLE
−80 −80

−100 −100
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Distance/km Distance/km

Figure 6. (a) The profile AA’ shown as brown dashed line in the map along with the source mechanisms from [Paul et al., 2015] in this
region. The mechanisms at depths ≤ 20 km are coloured grey while those of deeper events are coloured black. The earthquakes deeper
than 20 km have a dominant strike-slip component. The stations are represented by yellow triangles. (b) The CCP stack of receiver
functions (Gaussian width 2.5) along the profile AA’ overlain with inverted models (grey) for each cluster. The relative amplitude of
the phases w.r.t direct arrival are scaled. Positive amplitudes are coloured red and negative are in blue. Important structural features
like the PT, MCT, MHT, mid-crustal discontinuity are highlighted by dashed yellow lines. Oblique black lines identify the transverse
structures seen in the CCP. (c) A schematic diagram based on the inverted models and the CCP stack is shown for the region. The
MHT and other thrust sheets are represented by blue, the mid-crust in brown and the Moho in black dashed lines. The active fault
features (TS) are shown as cyan line-segments. The average VS in each of the region is also specified. (d) Comparative Moho depths
from H-K stacking results (green rectangles) and joint inversion results (blue triangles) with error bars. Data points from each of the
stations are arranged from the south to the north (based on the back-azimuth range of the cluster).

American Geophysical Union. All Rights Reserved.


American Geophysical Union. All Rights Reserved.
Viewed from ESE
(a) Viewed from top (c) 00 −8
27.9
−8 −10
5−5
−12
27.8
−10 10
−10
−14
Depth/km

15
−15 −16
27.7 −12
−18
MCT 20
−20
−20
−14
27.6
Latitude/degree

25
−25 −22
−16 −24
PT 30
−30
27.5
27.2 27.3 27.4 27.5 88.6 88.5 88.4
27.6 27.7 27.8 27.9 88.8 88.7
−18
Longitude/degree
27.4 OR
Viewed from NE
(d) 00
−20
−8
5−5
−10
27.3
OR −22
10
−10
−12
?
?

Depth/km
15
−15 −14
27.2 −24
−16
20
−20
88.35 88.4 88.45 88.5 88.55 88.6 88.65 88.7 88.75 88.8 −18
Longitude/degree 25
−25
−20
30
−30
−22
88.4
(b) 88.5
88.6 −24
M 88.7
88.8
? 27.2 27.3 27.4 27.5 27.6 27.7 27.8 27.9
Longitude/degree Latitude/degree
?
G
?
R

2017
?
N

c
Figure 7. 3-D plot of the MHT surface: (a) map view, (b) schematic illustrating the dome structure observed in map view, (c) view
from the E-SE direction, and (d) view from the NE direction. The MHT topography (depth 8–25 km) is colour-coded. The receiver
function piercing points are plotted as black asterisks and the stations are represented by white inverted triangles. The outcrop of the
PT and MCT are plotted on top. In map view, depth isolines are plotted as white (8–9 km) and blue (10–12 km) dashed lines. These
iso-surfaces are joined by cyan dashed lines, representing oblique ramps. The piercing points from NAMC, RABN, GTOK and MANG
are denoted as N, R, G, M, respectively, in the schematic plot (b). In (d), the dashed black lines join the MHT 3-D surface to the
outcrop. The thick magenta line represents the surface trace of the MHT locking line taken from Ader et al. [2012].
Figure 8. Map view of the Moho (depth 44–61 km) topography. The receiver function piercing

points are plotted as black asterisks and the stations are plotted as white triangles. The outcrops

of PT and MCT are plotted as black lines. Earthquake source mechanisms (depth ≥ 20 km)

from Paul et al. [2015] are overlay plotted.

2017
c American Geophysical Union. All Rights Reserved.

View publication stats

You might also like