zhang2021

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Earth and Planetary Science Letters 567 (2021) 116987

Contents lists available at ScienceDirect

Earth and Planetary Science Letters


www.elsevier.com/locate/epsl

Indian continental lithosphere and related volcanism beneath


Myanmar: Constraints from local earthquake tomography
Guangli Zhang a,c , Yumei He a,b,c,∗ , Yinshuang Ai a,b,c,∗∗ , Mingming Jiang a,b,c ,
Chit Thet Mon a,c,d , Guangbing Hou a , Myo Thant e,f , Kyaing Sein g
a
Key Laboratory of Earth and Planetary Physics, Institute of Geology and Geophysics, Chinese Academy of Sciences, 100029, Beijing, China
b
Innovation Academy for Earth Science, Chinese Academy of Sciences, 100029, Beijing, China
c
University of Chinese Academy of Sciences, 100049, Beijing, China
d
Department of Geology, Dagon University, 11422, Yangon, Myanmar
e
Department of Geology, Yangon University, 11041, Yangon, Myanmar
f
Myanmar Earthquake Committee, Hlaing Universities’ Campus, Hlaing Township, 11052, Yangon, Myanmar
g
Myanmar Geosciences Society, Yangon, Myanmar

a r t i c l e i n f o a b s t r a c t

Article history: The Indian plate descends obliquely eastward beneath the Eurasian plate along the Burmese arc. Previous
Received 5 November 2020 tomographic results revealed a high-velocity structure that plunges eastward into the deep mantle
Received in revised form 29 April 2021 beneath Myanmar. However, the shallow structure beneath Myanmar remains unclear due to the lack
Accepted 1 May 2021
of local seismic observations. Based on the local seismic data recorded by a newly deployed dense
Available online 13 May 2021
array in Myanmar, we obtain a three-dimensional velocity structure of the crust and lithospheric mantle
Editor: H. Thybo
above 100 km by employing the double-difference tomography method. Our imaging results support the
Keywords: existence of the Indian continental lithosphere to a depth of at least 100 km with a dip angle of ∼25◦
continental collision beneath the Indo-Burma Ranges and the Central Myanmar Basin. At the deep end of the continental slab,
Indian continental lithosphere the high V P , high V s and rather low average V P / V S ratio indicate that the lowermost portion of the
local earthquake tomography crust from 80 to 120 km may have experienced partially metamorphic eclogitization. The imaging results
Myanmar also provide direct seismic evidence for the origin of the last Monywa volcanic activity in the Holocene.
This subduction-related volcanism is characterized by a prominent low V s anomaly in the lithospheric
mantle, which indicates continental lithosphere dehydration and partial melting in mantle wedge. These
findings provide new seismic constraints to understand the continental collision system between India
and Eurasia under Myanmar and the related magmatic activities of the Monywa volcano.
© 2021 Elsevier B.V. All rights reserved.

1. Introduction Liang et al., 2016), whereas the existence of a continental litho-


sphere in the eastern Himalayan syntaxis is still controversial.
Myanmar, located at the southern edge of the eastern Hi-
Following the continent-continent collision between the Indian
malayan syntaxis, connects the ongoing continental collision along
and Eurasian plates beginning in the Eocene, the large-scale Indian
the Himalayas in the north to the oceanic subduction under An-
continental lithosphere subducted under Eurasia along its conver-
daman Sea in the south (Pesicek et al., 2010). In Myanmar, the
gent boundary (Capitanio et al., 2010; Replumaz et al., 2010; In-
distribution of intermediate-depth seismicity outlines an eastward
galls et al., 2016). In the western Himalayan syntaxis and central
Benioff zone extending down to a depth of ∼150 km (e.g., Ni et
Himalaya, the fronts of the subducted Indian continental litho-
al., 1989; Rao and Kalpna, 2005; Mon et al., 2020). However, the
sphere have been well imaged beneath the Hindu Kush and Ti-
seismic velocity structure along this region remains enigmatic. Pre-
betan Plateau using seismic tomography (e.g., Kufner et al., 2017;
vious teleseismic tomographic results have revealed a high-velocity
structure related to subducted plate reaching down to the deep
mantle but with very limited resolution at shallow depths (e.g., Li
* The second corresponding author. et al., 2008; Koulakov, 2011). Moreover, the regional-scale seismic
** The first corresponding author. Correspondence to: Key Laboratory of Earth and tomographic results have not yet provided unequivocal constraints
Planetary Physics, Institute of Geology and Geophysics, Chinese Academy of Sci-
ences, 100029, Beijing, China. on the detailed shape and nature of the subducted plate due to
E-mail addresses: ymhe@mail.igcas.ac.cn (Y. He), ysai@mail.iggcas.ac.cn (Y. Ai). the sparse seismic stations (Thiam et al., 2017; Raoof et al., 2017;

https://doi.org/10.1016/j.epsl.2021.116987
0012-821X/© 2021 Elsevier B.V. All rights reserved.
G. Zhang, Y. He, Y. Ai et al. Earth and Planetary Science Letters 567 (2021) 116987

Fig. 1. Simplified tectonic of the Myanmar region, showing major tectonic blocks, seismic stations and major faults. The main geological units are delimited by different
colors. In particular, the dark green color indicates the Wuntho-Monywa-Popa arc (WMPA), and the pink color delineates the Mogok Metamorphic Belt (MMB). The black
triangles show the seismic stations in the CMGSMO seismic network. The major faults plotted including the Churachandpur-Mao Fault (CMF), Kabaw Fault (KBF), and Sagaing
Fault (SGF), are plotted with north-south trending red lines. The west-east trending dotted blue lines and red line indicate the locations of the profiles shown in Fig. 7 and
Fig. 8, respectively. The inset map in the lower right corner shows the location of the study region in the Indo-Eurasia collision system. (For interpretation of the colors in
the figure(s), the reader is referred to the web version of this article.)

Wang et al., 2019). Zheng et al. (2020) detected the entire conti- 3-D lithospheric velocity structure of central Myanmar with the
nental crust along ∼22◦ N beneath Myanmar with receiver func- double-difference tomography method (Zhang and Thurber, 2003,
tion imaging; however, the three-dimensional velocity structure 2006). New local imaging results are able to characterize the high-
of the continental slab remains unclear, and the related tectonic resolution velocity structure of the crust and lithospheric mantle
regime and the origin of volcanism are also inconclusive. beneath Myanmar, which can help us better understand the sub-
In the context of plate collision and subduction, Myanmar can duction mechanism beneath Myanmar and the related volcanism
be mainly divided into four tectonic regimes from west to east: of the MWV.
the Indo-Burma Ranges (IBR), the Central Myanmar Basin (CMB),
the Mogok Metamorphic Belt (MMB), and the Shan Plateau (SP) 2. New dataset for local earthquake tomography
(Fig. 1). In the CMB, the Monywa volcano (MWV) occupies the cen-
ter of the north-south trending Wuntho-Monywa-Popa arc (WMPA, The CMGSMO seismic network consists of 71 portable broad-
in Fig. 1), which divides the CMB into a western forearc basin and band seismic stations (Fig. 1), with an average station spacing of
an eastern back-arc basin (Cai et al., 2019). The origin of the MWV ∼15 km for the main-line profile (38 stations, mainly along ∼22◦
is still under debate. Some models suggest a deep origin from the N) and ∼30 km for the other sporadic stations (33 stations) around
asthenosphere induced by slab rollback (Lee et al., 2016) or mantle the profile. From June 2016 to February 2018, the CMGSMO seis-
flow from a slab window followed by slab tear and “break-off” (Li mic network recorded more than one and a half years of continu-
et al., 2019; Zhang et al., 2020). Others prefer a subduction-related ous seismic data.
origin due to the dehydration of the subducted Indian lithosphere To build the seismic dataset for local earthquake tomography,
(Gardiner et al., 2015; Licht et al., 2020). However, direct structural the short-term average to long-term average (STA/LTA) algorithm
evidence from seismic imaging is still lacking, so a high-resolution is used to scan and identify local earthquakes from the continu-
local earthquake tomography study is urgently needed to map the ous waveform data. Here, we only select events that are recorded
lithospheric structure present beneath Myanmar to constrain the by eight or more stations to ensure the reliability of earthquake
origin of the MWV. detection, and then the first P- and S-arrivals of each event are
In 2016, under the China-Myanmar Geophysical Survey in the manually picked (Mon et al., 2020). All events included in the final
Myanmar Orogen (CMGSMO) project, the first portable seismic catalog are required to have at least six P-picks and three S-picks,
array was deployed in north-central Myanmar, with more than and the root mean square (RMS) residual of these picks is less than
one and a half years of seismic monitoring. In this study, based 3 s relative to the calculated travel time using CRUST 1.0 (Laske et
on the newly recorded dense array data, we imaged the detailed al., 2013). Using these criteria, we obtained 2316 local earthquakes,

2
G. Zhang, Y. He, Y. Ai et al. Earth and Planetary Science Letters 567 (2021) 116987

Fig. 2. The ray coverage in the study area from the map view shown in (a) and cross-section view shown in (b) and (c). The cyan dots indicate the earthquakes used for
inversion and the white triangles represent the seismic stations in the CMGSMO network. The dark blue lines show the ray path in the local earthquake tomography dataset.
(d) The number of earthquakes at different depths.

including 45674 P arrivals and 33308 S arrivals. Fig. 2 shows the s with an RMS value of 1.09 s, while those of the S-arrivals range
ray paths of these phases, in which we can see that the ray cov- from −5 s to 5 s and have an RMS value of 1.65 s.
erage inside the study area is sufficiently high at shallow depth. Regularization parameters including smoothing and damping
Though the ray coverage decreases with the increasing depth due values should be determined to find the optimum balance between
to the distribution of earthquakes, a considerable subset of earth- model roughness and data variance (Zhang and Thurber, 2003).
quakes is distributed in the depth range of 50–100 km (Fig. 2d), Here, we use the trade-off curve to obtain the optimal smooth-
which can help us to illuminate the subducted lithospheric struc- ing and damping value by trialing different combinations of these
ture beneath Myanmar. two parameters, and joint inversion of both event locations and
velocity structures is performed only once for each combination.
3. Tomographic details By analyzing the trade-off curve, the optimal damping value and
smoothing value are determined in the vicinity of the point with
Placing the picked first-arrivals into a time-distance axis, the P- the maximum curvature (Fig. 4).
and S-arrivals are linearly regressed to determine the average V P As the tomoDD method combines both the absolute and differ-
and V S (Fig. 3a). With this high-quality dataset, we perform local ential data between earthquake pairs, different weighting scenarios
earthquake tomography by utilizing the double-difference tomog- during inversion are employed for these two types of data to hi-
raphy algorithm (tomoDD). The velocity model is parameterized erarchically determine the velocity structure (Table S1). We tested
with 15 × 19 × 15 nodes in the west, north and depth directions different scenarios and found that a higher weight to the abso-
and smoothly interpolated with spline patches to form a contin- lute data at the beginning of inversion followed by adding more
uous velocity field. The vertical interval of the nodes is 10 km, weight to differential data in the later iterations has the lower RMS
whereas the horizontal node spacing is 26 km, which roughly values of 0.1958 s. This means that the large-scale velocity struc-
equals the average station spacing. The starting 1-D velocity model ture is first resolved by the absolute data, and then the detailed
for tomoDD inversion is derived from the VELEST program (Kissling structure near the source region is refined mainly by the differen-
et al., 1994). To obtain the optimal 1-D velocity model, different in- tial data. Thus, this method can obtain more accurate results near
put models were evaluated by calculating the travel time residuals the source region as the differential process reduces the picking
between the synthetic data and the real data (Fig. S1). Then, the errors and concentrates the ray-path derivatives near the source
1-D velocity model with the lowest RMS for both P- and S-arrivals region, which is particularly sensitive to the subducting structure
was obtained (Fig. 3b). In Figs. 3c–d, the perturbations between where there are more events inside and in the surrounding area
the P-arrivals and synthetic data range mainly from −2.5 s to 2.5 (Zhang et al., 2019). Compared to the imaging results of V P and

3
G. Zhang, Y. He, Y. Ai et al. Earth and Planetary Science Letters 567 (2021) 116987

Fig. 3. (a) The picked P- and S-arrivals. (b) The obtained 1-D velocity model by VELEST. (c–d) Time residuals between the synthetic data based on the 1-D velocity model and
the real data.

Fig. 4. Trade-off curve of the regularization parameters. The curvature of the trade-off curve is also shown with gray dashed lines (with arbitrary amplitude scaling). (a)
Relationship between the solution norm and residual norm across different damping situations is obtained with a fixed smoothing value of 20. The solution norm means
that both earthquake location and slowness variables are included in the norm calculation of the model variance. The optimal damping parameter of 150 is selected and
marked in red. (b) Relationship between the slowness norm and residual norm in different smoothing situations obtained with a fixed damping value of 150. The slowness
norm means that only the slowness is included for the norm calculation of the model variance. The optimal smoothing parameter of 20 is selected and marked in red. To be
more intuitive, the coordinate ranges in Fig. 4(a) and (b) have been normalized to the 0–1 range.

V S , the V P / V S ratio is more sensitive to the existence of fluids bility. Regularly alternating velocity anomalies with amplitudes of
(Takei, 2002). Thus, we inverted V P and V S separately with the ±10% relative to the initial 1-D velocity model shown in Fig. 3b,
same event phases and then divided the inverted results to obtain which are slightly less than the maximum amplitudes in the real
the V P / V S ratio to minimize the errors caused by different sizes data, are specified at each grid node as the “true” input. Gaussian
of P- and S-wave datasets (Figs. 7g–i). noise is also added to the synthetic travel times with a zero mean
To identify the relatively well-resolved areas in our model do- and standard deviations of 0.5 s for P-waves and 0.8 s for S-waves.
main, we calculate the derivative weight sum (DWS) values and Then, the synthetic data are inverted with the same inversion
perform the checkerboard test on both P- and S-wave velocity scheme and parameters as those used in real data inversion, in-
models. The DWS is a weighted operator representing the sum cluding the joint inversion of both event locations and slowness.
of partial differentials of travel time and slowness in grid nodes, The recovered results of the checkerboard tests on P-wave and
which indicates the density of ray sampling. The DWS values of S-wave data are shown in Figs. S5 and S6, respectively. The ro-
the P- and S-waves throughout the whole imaging area are shown bustness of our imaging results can be quantitatively measured by
in Fig. S3 and Fig. S4. In addition, the checkerboard resolution test computing the DWS values and analyzing the checkerboard test.
is also performed to provide another constraint for model relia- Based on these two evaluation indicators, the imaging areas where

4
G. Zhang, Y. He, Y. Ai et al. Earth and Planetary Science Letters 567 (2021) 116987

Fig. 5. Hypocenter distribution of the relocated earthquakes in 3-D view. Dots in different colors indicate earthquakes at different depth ranges. (a) and (b) represent different
perspectives. The black triangles show the seismic stations in the CMGSMO seismic network. The main faults, including the Churachandpur-Mao Fault (CMF), Kabaw Fault
(KBF), and Sagaing Fault (SGF), are shown in red lines.

the DWS is 100 or greater and the checkerboard recovery is greater ∼23◦ N through the Indaw anticline, which shows the detailed ge-
than 75% are outlined by the white dashed lines acting as the reli- ometry and layered structure of the forearc basin (Fig. 8a). The
able imaging zone (Figs. 6 and 7). However, checkerboard recovery thickness of the forearc basin, which is defined as 20–30 km in our
can be regarded as being on the optimistic side of the “truth” velocity models, agrees remarkably well with that of the basin in
because the same parametrization has been applied to the recov- the reflection profile (Fig. 8b). The extension along the west-east
ered model and the input model. Considering that the conventional direction of the forearc basin is also consistent with the reflec-
tightly spaced checkerboard may mislead the impression of struc- tion profile between the KBF and the WMPA. Moreover, the layered
tural distortion or smearing caused by data coverage (Rawlinson structure shown in Fig. 8a of the basin corresponds well with the
and Spakman, 2016), discrete spike tests that have the same scale velocity variations in our results.
length and amplitude of such anomalies were also performed and At 5 km depth, V P is relatively higher along the WMPA than
are shown in Figs. S7 and S8. in the neighboring low V P zones between the two basins, and
reaches the highest value of ∼6.5 km/s (Fig. 6a) beneath the MWV.
4. Results and interpretations At 35 km depth, a rather low V P anomaly of ∼7.5 km/s occurs be-
low the MWV, with a low V s anomaly of ∼4.2 km/s, hinting at the
4.1. Crustal velocity structure possible existence of arc magma. The absolute V s in the horizontal
slices, which are comparable to V P , can be seen in Fig. S9.
The absolute V P and V S models at crustal scale are displayed
in Figs. 6a–c and Fig. 7. In the upper-to-mid crust from 5 to 15 km, 4.2. Mantle velocity structure
the variation in V P is consistent with the shallow tectonic units of
central Myanmar marked by two evident tectonic boundaries. One In lithospheric mantle, a prominent north-south trending low
boundary is along the KBF, which exhibits the diving line of the V P belt is imaged beneath the IBR and CMB along the intermed-
low V P anomaly (3.5–4.0 km/s) and the high V P anomaly (5.5–6.0 iate-depth seismicity (Figs. 6d–f). This low V P velocity belt (∼7.5
km/s), corresponding to the eastern thick sedimentary layer be- km/s) exists at depths from 35 to 85 km and gradually moves east-
neath the CMB and the western basement rocks beneath the IBR. ward as the depth increases. In Fig. 7, this low V P anomaly (LV1)
The other boundary is the WMPA, with a strong anomaly (6.0–6.5 is projected as an ∼30 km thick slab extending to a depth of ∼85
km/s) below this volcanic arc separating the CMB into two north- km with a dip angle of ∼25◦ in the vertical profiles. A slight de-
south trending low velocity zones (Figs. 6a–b). These two low V P crease in the thickness of LV1 can also be observed from north to
anomalous zones, associated with the low V S in Fig. S9 and the south. The V P /V s ratio is rather high (1.8–1.9) along this low V P
high V P / V S ratios in Fig. 7, correspond to the forearc basin and anomaly, whereas a significantly low V P / V S zone is found along
the back-arc basin in Myanmar. the intermediate-depth seismicity at the lower boundary of LV1
The low V P anomaly beneath the forearc basin is 20–30 km (inside the blue dotted rectangles in Figs. 7g–i), where V P (∼8.2
thick, much thicker than the low V P anomaly (10–15 km) beneath km/s) and V S (∼4.6 km/s) are slightly higher than LV1 but similar
the back-arc basin. Lower V S and higher V P / V S ratios are also ob- to the surrounding lithospheric mantle.
served beneath the forearc basin. To further validate the structure Starting at a depth of 55 km, a high V P anomaly (∼8.5 km/s)
of the forearc basin in our velocity model, we compare our results accompanies the LV1 anomaly on the west side and moves east-
at shallow depths with the previous seismic reflection profile in ward as the depth increases (Figs. 6d–f). This high V P anomaly
the Chindwin Basin (Böker et al., 2019). This profile is located be- (HV) dips at the same angle as the LV1 anomaly and extends to
tween the western margin of the Burma microplate to the west deeper depths (>100 km). The couple of LV1 and HV indicates
and the WMPA to the east, with a 120 km long extension along that they may be an intact slab beneath Myanmar (Fig. 7). More-

5
G. Zhang, Y. He, Y. Ai et al. Earth and Planetary Science Letters 567 (2021) 116987

Fig. 6. Absolute velocity of the P-wave in the horizontal slices at the 5 km, 15 km, 35 km, 55 km, 65 km and 85 km depth levels. The main faults including the Churachandpur-
Mao Fault (CMF), Kabaw Fault, and Sagaing Fault, are shown as black lines. The dashed black lines delineate the Wutho-Monywa-Popa arc (WMPA). The black drawn triangles
are seismic stations, and the gray dots indicate earthquakes located at the imaging depth level within the range of ±5 km. The marked volcanoes are the Monywa volcano
and Singu volcano. The areas inside the white dashed lines are the relatively reliable imaging areas we sketched.

over, a remarkably low V P anomaly (LV2) appears below the MWV In Figs. 9c–g, we tested five different end-member slab models
at depths ranging from 35 to 100 km in the CC’ profile. This low by varying the penetrating depth (80/120 km), breaking-off point,
V P anomaly (7.5–8 km/s) exists on the east side of the LV1 and HV thickness (30/10 km) of LV1 and the existence of LV2. The syn-
and connects with the low V P anomaly at a depth of 35 km. thetic travel-time data were first calculated based on these input
models and then used for restoration inversion with the same in-
5. Discussion version scheme as the real data. Gaussian random noise was also
added to each selected arrival with a mean of zero and a stan-
dard deviation of 0.5 s. The synthetic output results are plotted
5.1. Resolution test of the imaging results
in the right column of Figs. 9c–g. In the crust, all the velocity
structures can be well recovered. In the lithospheric mantle, the
The reliability of velocity anomalies in Section 4 can be eval- recovered LV1 anomaly appears as a continuous low-velocity zone
uated by resolution test. In this section, the absolute V P imaging with a well recovered geometry (Figs. 9c and 9g), which is similar
results along profile CC’ in Fig. 7c are chosen as the real abso- to the imaging results shown in Fig. 9a. Even when LV1 penetrates
lute velocity model (Fig. 9a), and the relative V P results compared deeper to a depth of 120 km (Fig. 9d), it can be clearly illumi-
to the initial velocity model (Fig. 3b) are set as the real relative nated inside the resolution limit. A breaking-off point has also
velocity model (Fig. 9b). To create the synthetic input model, we been included in the slab model at depths ranging from 50–70
directly used the 1-D velocity in Fig. 3b as the crustal input model km where earthquakes are poorly distributed, and the break of the
and mimicked the geometry, penetrating depth, and intensity of slab can be clearly seen in the recovered LV1 anomaly (Fig. 9e).
the main velocity anomalies in the lithospheric mantle (i.e., LV1, The model with an LV1 thickness of 10 km (Fig. 9f) can be no-
HV, and LV2) in Fig. 9b. To minimize the impact of precondition- tably distinguished from the models with an LV1 thickness of 30
ing, a finer-scale parametrization with 2 × 2 km node spacing is km (Fig. 9c–e) with poor LV1 recovery. Moreover, the recovered HV
applied to the input model, which allows us to construct veloc- anomaly beneath LV1 is slightly vague at the edge of the resolu-
ity anomalies much finer than the inversion grid (Rawlinson and tion limit due to the insufficient distribution of ray paths below
Spakman, 2016). Thus, LV1 is set as an ∼30 km thick anomaly with it.
a V P of 7.5 km/s, while the HV anomaly has a thickness of >80 Comparing the synthetic output with the real velocity model
km with a V P of 8.7 km/s and the same dip angle of ∼25◦ be- (Fig. 9a), the anomalies in LV1, HV and LV2 are shown to be re-
neath it (Fig. 9c). The LV2 anomaly with a V P of 7.5 km/s (Fig. 9g) liable. The LV1 anomaly (∼30 km thick and V P of 7.2–8 km/s)
simulates the upwelling geometry in Fig. 7c beneath the MWV. extends to a depth of 80 km in the lithospheric mantle beneath

6
G. Zhang, Y. He, Y. Ai et al. Earth and Planetary Science Letters 567 (2021) 116987

Fig. 7. Imaging results of V P , V S , and the V P /V s ratio in three vertical profiles. The vertical profiles start from 23◦ N to 22◦ N at an interval of 0.5◦ , totaling three east-west
trending vertical profiles marked as profiles AA’, BB’, and CC’ (see Fig. 1 for locations). (a–c) Absolute V P results in the vertical profiles. (d–f) Absolute V S results in the
vertical profiles. (g–i) The V P /V s ratio in the vertical profiles. The gray dots indicate the earthquakes relocated at the imaging latitude within the range of ± 0.125◦ . The
areas inside the white dashed lines are the relatively reliable imaging areas we sketched. The dashed black lines indicate the Moho depths obtained from CRUST 1.0 (Laske
et al., 2013). IBR: Indo-Burma Ranges; KBF: Kabaw Fault; SGF: Sagaing Fault; MWV: Monywa volcano.

Fig. 8. Comparison between our results and the seismic reflection profile. The seismic reflection profile in (a) is modified from Böker et al. (2019). The location of the profile
is given as a red line in Fig. 1. Subfigure (b) shows the S-wave velocity model beneath the forearc basins at the same location along 23◦ N between the KBF and the WMPA.
The black lines represent contours of the S-wave velocity.

7
G. Zhang, Y. He, Y. Ai et al. Earth and Planetary Science Letters 567 (2021) 116987

Myanmar, and the HV anomaly is present beneath it with the same of LV1 from 80 to 120 km (inside the blue dotted rectangles in
dipping angle of ∼25◦ . There also exists an LV2 anomaly beneath Figs. 7g–i), where the V P (∼8.2 km/s) and V s (∼4.6 km/s) are
the MWV with a lower V P of ∼7.5 km/s. Similar resolution tests higher than those of LV1 (V P of 7.2∼7.8 km/s; V s of 4.0–4.5
of V S imaging results are shown in Fig. S10. km/s) but similar to those of the surrounding lithospheric mantle.
The average V P / V S ratios, which are more sensitive to the ex-
5.2. Indian continental lithosphere beneath Myanmar istence of fluids, have also been noted to be rather low (∼1.68)
within this depth range along the seismicity (Figs. 7g–i). Com-
In our imaging results, there is an eastward-dipping low- pared to the results of seismological investigations on the sub-
velocity anomaly LV1 with a thickness of ∼30 km and a dip angle ducted lower crust beneath southern Tibet and Pamir, such V P and
of ∼25◦ extending to a depth of 80 km beneath the IBR and the V S values in the lower crust are typical indicators of the partial
CMB (Fig. 7). The morphology and velocity characteristics (∼30 eclogitization of subducted continental lower crust (e.g., Schulte-
km thickness, V P ∼7.5 km/s, and V S ∼4.2 km/s) of the LV1 indi- Pelkum et al., 2005; Wittlinger et al., 2009; Sippl et al., 2013).
cate that it is continental crust, which is highly consistent with The localized eclogitization of the lower crust can be catalyzed by
previous receiver function imaging results (Zheng et al., 2020). dehydration-produced water and considerably reduces the strength
The thickness of this crustal layer in the mantle fits well with of rocks so that they are brittle in character (Kita et al., 2010;
the northeastern Indian crust (Islam et al., 2011; Bora and Baruah, Shi et al., 2020). This reaction-induced embrittlement may trigger
2012). Moreover, the V P values slightly above 7.2 km/s and V S val- intermediate-depth events in the lower continental crust (Incel et
ues slightly above 4.0 km/s at depths ≥35 km may also come from al., 2019), which provides a possible scenario for interpreting the
rocks with compositions typical of the northeast Indian crust in observed deep seismicity in our results.
the Bengal Basin (Mitra et al., 2018). Furthermore, a high-velocity In the metamorphic eclogitization process, the density of the
anomaly HV with the same dip angle as LV1 is found beneath the mafic continental lower crust can be increased to a value simi-
crust. Intermediate depth seismicity occurs at depths up to 150 lar to that of the lithospheric mantle (Hetényi et al., 2007), thus
km below Myanmar (Fig. 5), and these earthquakes are thought to reversing buoyancy and facilitating deep burial (Krystopowicz and
correspond well with LV1 and HV. Although the thickness of HV is Currie, 2013). However, water in the continental crust is scarcer
uncertain, it is still reasonable to interpret the LV1 and HV anoma- than that in the oceanic crust, so eclogitization may be slower
lies as the Indian continental lithosphere according to their intact or incomplete (Leech, 2001; Schneider et al., 2013). Such partial
morphology and the seismicity along them (Fig. 5). eclogitization may explain why the lower crust in our results can
As the continental lithosphere is generally considered too buoy- penetrate to at least 100 km beneath Myanmar without delamina-
ant to actively drive subduction, an external force is required to tion, whereas the subducted oceanic crust weakened by complete
explain the current continental lithosphere beneath Myanmar. Pre- eclogitization is usually decoupled from the lithospheric mantle at
vious teleseismic tomography has revealed a high-velocity anomaly shallow depths (Doin and Henry, 2001).
in the deep mantle dipping at ∼60◦ , which is interpreted as the
subducted Indian oceanic lithosphere (e.g., Li et al., 2008; Koulakov,
5.4. Holocene volcanism of the Monywa volcano (MWV)
2011). Although the continental lithosphere may be less dense
than the underlying mantle, it can be triggered by the attachment
of a denser oceanic slab after ocean closure (Capitanio et al., 2010). The MWV is located in the central part of the north-south
Thus, we conclude that the Indian continental lithosphere may ex- trending WMPA, which separates the CMB into forearc basin and
tend to a depth of at least 100 km beneath Myanmar and may back-arc basin (Wang et al., 2014; Cai et al., 2019). Previous geo-
have been pulled down by earlier oceanic subduction. chemical analysis indicates that the latest eruption of the MWV
In Fig. 6 and Fig. 7, a slight decrease in the thickness of LV1 occurred in the Holocene (Belousov et al., 2018), but the deep ori-
can be observed from north to south, which may indicate the pos- gin of this volcanism remains unclear due to the lack of seismic
sible transition from continental Indian lithosphere to transitional- observations. In the crust, a relatively high velocity anomaly (V P of
oceanic lithosphere in further south (Mallick et al., 2020). Com- 6.5 km/s; V s of 3.5 km/s) is observed beneath the MWV (Figs. 7c
bined with the deep oceanic slab revealed by previous imaging and 7f). Directly beneath it, a low velocity anomaly with V P of
results (e.g., Li et al., 2008; Koulakov, 2011), a three-dimensional 7.2 km/s and low V s of 4.2 km/s (LV2) is imaged, representing the
continent-to-ocean transition may exist in this subduction zone. As partially melting materials (Nakajima et al., 2001). Resolution tests
the subduction below the IBR is still ongoing with an ∼18 mm/yr have validated the shape and amplitude of these two anomalies
convergence rate (Steckler et al., 2016), the continued difference in (Fig. 9g). Here, we interpret these two anomalies as direct struc-
the buoyancies of the continental and oceanic lithosphere in this tural evidence of the continental subduction-related volcanism of
transition zone may trigger slab tearing and break-off (Zhang et al., the MWV.
2020). This may occur at a deeper depth as the intermediate depth The volcanic rocks from the latest volcanism of the MWV
seismicity is distributed above a depth of ∼150 km (Fig. 5). Thus, are featured as calc-alkaline and shoshonitic affinities exhibiting
the Indian continental lithosphere beneath Myanmar may also be subduction-related signatures (Lee et al., 2016; Zhang et al., 2020).
a result of rebound or uplift after slab detachment (Wortel and Based on the imaging results shown in Fig. 7c, we speculate that
Spakman, 2000). However, more seismic evidence of slab break-off the Indian continental lithosphere extends eastward just below
and tearing beneath Myanmar is needed at mantle depths. the MWV, where the LV2 anomaly delineates a mantle upwelling
process. This indicates that fluid was released from the cold conti-
5.3. Partial eclogitization of the subducted Indian lower crust nental lithosphere and migrated into the overlying mantle wedge
at depths from 80 to 120 km, resulting in the partial melting of
The intermediate-depth events shown in Fig. 7 are relocated si- the mantle wedge and the formation of calc-alkaline arc magma
multaneously with the inversion of the velocity structure (Zhang (Lee et al., 2016). At depths ranging from 30 to 40 km, the LV2
and Thurber, 2003). The mean location errors in the latitude, longi- anomaly acts as a hanging pocket with a lower V P of 7.2 km/s and
tude and depth of these earthquakes are 3.16 km, 3.98 km and 4.31 a lower V s of 4.0 km/s (Figs. 7c and 7f), indicating the possible
km, respectively (Fig. 5). Most of the intermediate-depth events are existence of a magma chamber filled with upwelling melting ma-
distributed along the lowermost part of LV1 at depths of 40–80 terials (Koulakov et al., 2009). Then the magma chamber may feed
km. However, the seismicity extends beyond the down deep end the surface volcanism and hydrothermal activities of the MWV, and

8
G. Zhang, Y. He, Y. Ai et al. Earth and Planetary Science Letters 567 (2021) 116987

Fig. 9. Synthetic tests performed to evaluate the resolution capability of the local earthquake tomography with respect to the three velocity anomalies in the mantle. (a) Real
absolute velocity along profile CC’ in Fig. 7c. Contours are given in 0.5 km/s intervals. The white dashed lines depict the resolution limit. The dashed black lines indicate
the Moho depths obtained from CRUST 1.0 (Laske et al., 2013). (b) Real relative velocity compared to the initial velocity model along profile CC’. (c–g) Five end member
tests performed to validate the penetrating depth (80/120 km), the thickness of LV1 (30/10 km), the break-off point of the slab and the existence of LV2. The left column
shows the “SYN INPUT” which is the synthetic input model constrained by mimicking the results from (a) and (b). The right column shows the “SYN OUTPUT”, which is the
inversion result of the synthetic input. The gray dots indicate the earthquakes located between latitudes 21.875◦ N and 22.125◦ N. IBR: Indo-Burma Ranges; KBF: Kabaw Fault;
SGF: Sagaing Fault; MWV: Monywa volcano.

9
G. Zhang, Y. He, Y. Ai et al. Earth and Planetary Science Letters 567 (2021) 116987

Fig. 10. An illustration of our interpretation of the crustal and lithospheric mantle structure beneath Myanmar. Contours represent 0.5 km/s intervals. The white dashed lines
depict the resolution limit. The dashed black lines indicate the Moho depths obtained from CRUST 1.0 (Laske et al., 2013). The gray dots denote the relocated shallow and
intermediate-depth earthquakes. IBR: Indo-Burma Ranges; KBF: Kabaw Fault; WMPA: Wuntho-Monywa-Popa arc; MWV: Monywa volcano.

a low V P / V S ratio in the crust may indicate the presence of water doi.org/10.12129/IGGSL.Data.Observation), for their efforts to in-
(Takei, 2002). stall and maintain the CMGSMO seismic network. We thank Xin
Wang for the linguistic assistance during the preparation of the
6. Conclusions manuscript. Plots were generated using Generic Mapping Tools
(GMT). This work was supported by the National Natural Science
In this study, we have developed high-resolution V P , V S , and Foundation of China (grants 91755214, 42030309, 41490612), the
V P / V S models of the crust and lithospheric mantle beneath cen- Strategic Priority Research Program (B) of the Chinese Academy
tral Myanmar, which provide new constraints on the existence of of Sciences (XDB18000000) and the International Partnership Pro-
Indian continental lithosphere beneath Myanmar and the origin of gram of the Chinese Academy of Sciences (GJHZ1776).
the Monywa volcano. The Indian continental lithosphere beneath
the IBR and CMB is well imaged in our tomographic model. These Appendix A. Supplementary material
results reveal the geometries and characteristics of the continental
regime extending to a depth of at least 100 km with a dip angle of Supplementary material related to this article can be found on-
∼25◦ (Fig. 10). This continental lithosphere beneath Myanmar may line at https://doi.org/10.1016/j.epsl.2021.116987.
have been pulled down by the attachment of the earlier subducted
Indian oceanic lithosphere. References
At the down deep end of the Indian continental slab, the dis-
Belousov, A., Belousova, M., Zaw, K., Streck, M.J., Bindeman, I., Meffre, S., Vasconce-
tribution of intermediate-depth seismicity, with a high V P of 8.2 los, P., 2018. Holocene eruptions of Mt. Popa, Myanmar: volcanological evidence
km/s, a high V s of 4.6 km/s and a rather low average V P / V S ratio of the ongoing subduction of Indian Plate along Arakan Trench. J. Volcanol.
of ∼1.68, indicates that the subducted Indian lowermost crust may Geotherm. Res. 360, 126–138. https://doi.org/10.1016/j.jvolgeores.2018.06.010.
have experienced partial eclogitization, which provides a possible Böker, U., Leo, C., Thu, P.S., 2019. Structural and stratigraphic review of the South-
ern Chindwin Basin, Myanmar. In: Myanmar Petroleum Technology Conference.
scenario to interpret the deep seismicity in our results.
Yangon, Myanmar.
Our velocity models also provide direct seismic structure evi- Bora, D.K., Baruah, S., 2012. Mapping the crustal thickness in Shillong-Mikir Hills
dence of the origin of the volcanism of the MWV. The low V P and Plateau and its adjoining region of northeastern India using Moho reflected
V s anomalies in the lithospheric mantle, indicate the continental waves. J. Asian Earth Sci. 48, 83–92. https://doi.org/10.1016/j.jseaes.2011.12.007.
Cai, F., Ding, L., Zhang, Q., Orme, D.A., Wei, H., Li, J., et al., 2019. Initiation and
lithosphere dehydration and partial melting in the mantle wedge.
evolution of forearc basins in the central Myanmar depression. Geol. Soc. Am.
At depths of 30–40 km, the imaged low velocity zone indicates Bull. https://doi.org/10.1130/B35301.1.
the possible existence of a magma chamber, which may be filled Capitanio, F.A., Morra, G., Goes, S., Weinberg, R.F., Moresi, L., 2010. India–Asia
by upwelling materials from continental subduction and feed the convergence driven by the subduction of the Greater Indian continent. Nat.
surface volcanism and hydrothermal activities of the MWV. Geosci. 3, 136–139. https://doi.org/10.1038/ngeo725.
Doin, M.P., Henry, P., 2001. Subduction initiation and continental crust recycling:
the roles of rheology and eclogitization. Tectonophysics 342, 163–191. https://
Declaration of competing interest doi.org/10.1016/S0040-1951(01)00161-5.
Gardiner, N.J., Searle, M.P., Robb, L.J., Morley, C.K., 2015. Neo-Tethyan magmatism
and metallogeny in Myanmar - an Andean analogue? J. Asian Earth Sci. 106,
The authors declare that they have no known competing finan-
197–215. https://doi.org/10.1016/j.jseaes.2015.03.015.
cial interests or personal relationships that could have appeared to Hetényi, G., Cattin, R., Brunet, F., Bollinger, L., Vergne, J., Nábělek, J.L., Diament,
influence the work reported in this paper. M., 2007. Density distribution of the India plate beneath the Tibetan Plateau:
geophysical and petrological constraints on the kinetics of lower-crustal eclogi-
tization. Earth Planet. Sci. Lett. 264, 226–244. https://doi.org/10.1016/j.epsl.2007.
Acknowledgements
09.036.
Incel, S., Labrousse, L., Hilairet, N., John, T., Gasc, J., Shi, F., Wang, Y., Andersen,
We thank Editor Hans Thybo and two anonymous reviewers T.B., Renard, F., Jamtveit, B., Schubnel, A., 2019. Reaction-induced embrittlement
for their constructive suggestions and comments that helped us of the lower continental crust. Geology 47, 235–238. https://doi.org/10.1130/
G45527.1.
significantly improve the paper. We gratefully acknowledge the re- Ingalls, M., Rowley, D.B., Currie, B., Colman, A.S., 2016. Large-scale subduction of
searchers and colleagues at the Seismic Array Laboratory, Institute continental crust implied by India-Asia mass-balance calculation. Nat. Geosci. 9,
of Geology and Geophysics, Chinese Academy of Sciences (http:// 848–853. https://doi.org/10.1038/ngeo2806.

10
G. Zhang, Y. He, Y. Ai et al. Earth and Planetary Science Letters 567 (2021) 116987

Islam, M.S., Shinjo, R., Kayal, J.R., 2011. Pop-up tectonics of the shillong plateau Pesicek, J.D., Thurber, C.H., Widiyantoro, S., Zhang, H., DeShon, H.R., Engdahl, E.R.,
in northeastern India: insight from numerical simulations. Gondwana Res. 20, 2010. Sharpening the tomographic image of the subducting slab below Sumatra,
395–404. https://doi.org/10.1016/j.gr.2010.11.007. the Andaman Islands and Burma. Geophys. J. Int. 182, 433–453. https://doi.org/
Kissling, E., Ellsworth, W.L., Eberhart-Phillips, D., Kradolfer, U., 1994. Initial refer- 10.1111/j.1365-246X.2010.04630.x.
ence models in local earthquake tomography. J. Geophys. Res. Solid Earth. 99, Rao, N.P., Kalpna, 2005. Deformation of the subducted Indian lithospheric slab in the
19635–19646. https://doi.org/10.1029/93JB03138. Burmese arc. Geophys. Res. Lett. 32, 1–5. https://doi.org/10.1029/2004GL022034.
Kita, S., Okada, T., Hasegawa, A., Nakajima, J., Matsuzawa, T., 2010. Anomalous deep- Raoof, J., Mukhopadhyay, S., Koulakov, I., Kayal, J.R., 2017. 3-D seismic tomography
ening of a seismic belt in the upper-plane of the double seismic zone in the of the lithosphere and its geodynamic implications beneath the northeast India
Pacific slab beneath the Hokkaido corner: possible evidence for thermal shield- region. Tectonics 36, 962–980. https://doi.org/10.1002/2016TC004375.
ing caused by subducted forearc crust materials. Earth Planet. Sci. Lett. 290, Rawlinson, N., Spakman, W., 2016. On the use of sensitivity tests in seismic tomog-
415–426. https://doi.org/10.1016/j.epsl.2009.12.038. raphy. Geophys. J. Int. 205, 1221–1243. https://doi.org/10.1093/gji/ggw084.
Koulakov, I., 2011. High-frequency P and S velocity anomalies in the upper mantle Replumaz, A., Negredo, A.M., Guillot, S., Villaseñor, A., 2010. Multiple episodes of
beneath Asia from inversion of worldwide traveltime data. J. Geophys. Res. 116, continental subduction during India/Asia convergence: insight from seismic to-
B04301. https://doi.org/10.1029/2010JB007938. mography and tectonic reconstruction. Tectonophysics 483, 125–134. https://
Koulakov, I., Yudistira, T., Luehr, B.G., Wandono, 2009. P, S velocity and VP/VS ratio doi.org/10.1016/j.tecto.2009.10.007.
beneath the Toba caldera complex (Northern Sumatra) from local earthquake to- Schneider, F.M., Yuan, X., Schurr, B., Mechie, J., Sippl, C., Haberland, C., 2013. Seismic
mography. Geophys. J. Int. 177, 1121–1139. https://doi.org/10.1111/j.1365-246X. imaging of subducting continental lower crust beneath the Pamir. Earth Planet.
2009.04114.x. Sci. Lett. 375, 101–112. https://doi.org/10.1016/j.epsl.2013.05.015.
Krystopowicz, N.J., Currie, C.A., 2013. Crustal eclogitization and lithosphere delami- Schulte-Pelkum, V., Monsalve, G., Sheehan, A., Pandey, M.R., Sapkota, S., Bilham,
nation in orogens. Earth Planet. Sci. Lett. 361, 195–207. https://doi.org/10.1016/ R., Wu, F., 2005. Imaging the Indian subcontinent beneath the Himalaya. Na-
j.epsl.2012.09.056. ture 435, 1222–1225. https://doi.org/10.1038/nature03678.
Kufner, S.K., Schurr, B., Haberland, C., Zhang, Y., Saul, J., Ischuk, A., Oimahmadov, Shi, D., Klemperer, S.L., Shi, J., Wu, Z., Zhao, W., 2020. Localized foundering of Indian
I., 2017. Zooming into the Hindu Kush slab break-off: a rare glimpse on the lower crust in the India-Tibet collision zone. Proc. Natl. Acad. Sci. U. S. A. 117,
terminal stage of subduction. Earth Planet. Sci. Lett. 461, 127–140. https://doi. 24742–24747. https://doi.org/10.1073/pnas.2000015117.
org/10.1016/j.epsl.2016.12.043. Sippl, C., Schurr, B., Tympel, J., Angiboust, S., Mechie, J., Yuan, X., Schneider, F.M.,
Laske, G., Masters, G., Ma, Z., et al., 2013. Update on CRUST1.0-A 1-degree global Sobolev, S.V., Ratschbacher, L., Haberland, C., 2013. Deep burial of Asian conti-
model of Earth’s crust. Geophys. Res. Abstracts, Abstract EGU2013–2658. http:// nental crust beneath the Pamir imaged with local earthquake tomography. Earth
igppweb.ucsd.edu/~gabi/rem.html. Planet. Sci. Lett. 384, 165–177. https://doi.org/10.1016/j.epsl.2013.10.013.
Lee, H.Y., Chung, S.L., Yang, H.M., 2016. Late Cenozoic volcanism in central Myanmar: Steckler, M.S., Mondal, D.R., Akhter, S.H., Seeber, L., Feng, L., Gale, J., Hill, E.M.,
geochemical characteristics and geodynamic significance. Lithos 245, 174–190. Howe, M., 2016. Locked and loading megathrust linked to active subduction be-
https://doi.org/10.1016/j.lithos.2015.09.018. neath the Indo-Burman ranges. Nat. Geosci. 9, 615–618. https://doi.org/10.1038/
Leech, M.L., 2001. Arrested orogenic development: eclogitization, delamination, and ngeo2760.
tectonic collapse. Earth Planet. Sci. Lett. 185, 149–159. https://doi.org/10.1016/ Takei, Y., 2002. Effect of pore geometry on V P / V S : from equilibrium geometry to
S0012-821X(00)00374-5. crack. J. Geophys. Res. 107. https://doi.org/10.1029/2001jb000522.
Li, C., Van der Hilst, R.D., Meltzer, A.S., Engdahl, E.R., 2008. Subduction of the Indian Thiam, H.N., Htwe, Y.M.M., Kyaw, T.L., Tun, P.P., Min, Z., Htwe, S.H., et al., 2017. A re-
lithosphere beneath the Tibetan Plateau and Burma. Earth Planet. Sci. Lett. 274 port on upgraded seismic monitoring stations in Myanmar: station performance
(1–2), 157–168. https://doi.org/10.1016/j.epsl.2008.07.016. and site response. Seismol. Res. Lett. 88 (3), 926–934. https://doi.org/10.1785/
Li, J.X., Fan, W.M., Zhang, L.Y., Ding, L., Sun, Y.L., Peng, T.P., Cai, F.L., Guan, Q.Y., Sein, 0220160168.
K., 2019. Subduction of Indian continental lithosphere constrained by Eocene- Wang, X., Wei, S., Wang, Y., Maung Maung, P., Hubbard, J., Banerjee, P., et al., 2019.
Oligocene magmatism in northern Myanmar. Lithos 348–349, 105211. https:// A 3-D shear wave velocity model for Myanmar region. J. Geophys. Res., Solid
doi.org/10.1016/j.lithos.2019.105211. Earth 124, 504–526. https://doi.org/10.1029/2018JB016622.
Liang, X., Chen, Y., Tian, X., Chen, Y.J., Ni, J., Gallegos, A., Klemperer, S.L., Wang, M., Wang, Y., Sieh, K., Tun, S.T., Lai, K.Y., Myint, T., 2014. Active tectonics and earthquake
Xu, T., Sun, C., Si, S., Lan, H., Teng, J., 2016. 3D imaging of subducting and potential of the Myanmar region. J. Geophys. Res., Solid Earth 119, 3767–3822.
fragmenting Indian continental lithosphere beneath southern and central Ti- https://doi.org/10.1002/2013JB010762.
bet using body-wave finite-frequency tomography. Earth Planet. Sci. Lett. 443, Wittlinger, G., Farra, V., Hetényi, G., Vergne, J., Nábělek, J., 2009. Seismic velocities
162–175. https://doi.org/10.1016/j.epsl.2016.03.029. in southern Tibet lower crust: a receiver function approach for eclogite detec-
Licht, A., Win, Z., Westerweel, J., Cogné, N., Morley, C.K., Chantraprasert, S., Poblete, tion. Geophys. J. Int. 177, 1037–1049. https://doi.org/10.1111/j.1365-246X.2008.
F., Ugrai, T., Nelson, B., Aung, D.W., Dupont-Nivet, G., 2020. Magmatic history 04084.x.
of central Myanmar and implications for the evolution of the Burma Terrane. Wortel, M.J.R., Spakman, W., 2000. Subduction and slab detachment in the
Gondwana Res. 87, 303–319. https://doi.org/10.1016/j.gr.2020.06.016. Mediterranean-Carpathian region. Science 80 (290), 1910–1917. https://doi.org/
Mallick, R., Hubbard, J.A., Lindsey, E.O., Bradley, K.E., Moore, J.D.P., Ahsan, A., Khor- 10.1126/science.290.5498.1910.
shed Alam, A.K.M., Hill, E.M., 2020. Subduction initiation and the rise of the Zhang, H., Thurber, C.H., 2003. Double-difference tomography: the method and
Shillong Plateau. Earth Planet. Sci. Lett. 543, 116351. https://doi.org/10.1016/j. its application to the Hayward Fault, California. Bull. Seismol. Soc. Am. 93,
epsl.2020.116351. 1875–1889. https://doi.org/10.1785/0120020190.
Mitra, S., Priestley, K.F., Borah, K., Gaur, V.K., 2018. Crustal structure and evolution of Zhang, H., Thurber, C.H., 2006. Development and applications of double-difference
the Eastern Himalayan Plate boundary system, Northeast India. J. Geophys. Res., tomography. Pure Appl. Geophys. 163, 373–403. https://doi.org/10.1007/s00024-
Solid Earth 123, 621–640. https://doi.org/10.1002/2017JB014714. 005-0021-y.
Mon, C.T., Gong, X., Wen, Y., Jiang, M., Chen, Q.F., Zhang, M., et al., 2020. In- Zhang, H., Wang, F., Myhill, R., Guo, H., 2019. Slab morphology and deformation
sight into major active faults in central Myanmar and the related geodynamic beneath Izu-Bonin. Nat. Commun. https://doi.org/10.1038/s41467-019-09279-7.
sources. Geophysical Research Letters 47, e2019GL086236. https://doi.org/10. Zhang, L.Y., Fan, W.M., Ding, L., Ducea, M.N., Pullen, A., Li, J.X., Sun, Y.L., Yue, Y.H.,
1029/2019GL086236. Cai, F.L., Wang, C., Peng, T.P., Sein, K., 2020. Quaternary volcanism in Myanmar:
Nakajima, J., Matsuzawa, T., Hasegawa, A., Zhao, D., 2001. Three-dimensional struc- a record of Indian slab tearing in a transition zone from oceanic to continen-
ture of Vp, Vs and Vp/Vs Beneath Northeastern Japan arc: implications for arc tal subduction. Geochemistry, Geophys. Geosystems, 0. https://doi.org/10.1029/
magmatism and fluids. J. Geophys. Res. 106, 21843–21857. https://doi.org/10. 2020gc009091.
1029/2000JB000008. Zheng, T., He, Y., Ding, L., et al., 2020. Direct structural evidence of Indian continen-
Ni, J.F., Guzman-Speziale, M., Bevis, M., Holt, W.E., Wallace, T.C., Seager, W.R., 1989. tal subduction beneath Myanmar. Nat Commun 11, 1944. https://doi-org-443.
Accretionary tectonics of Burma and the three-dimensional geometry of the webvpn.las.ac.cn/10.1038/s41467-020-15746-3.
Burma subduction zone. Geology 17 (1), 68–71. https://doi.org/10.1130/0091-
7613(1989)017<0068:ATOBAT>2.3.CO;2.

11

You might also like