You are on page 1of 14

PUBLICATIONS

Journal of Geophysical Research: Solid Earth


RESEARCH ARTICLE Upper mantle discontinuity structure beneath eastern
10.1002/2016JB013551
and southeastern Tibet: New constraints
Key Points:
Topographic variation on the two
on the Tengchong intraplate volcano
discontinuities across the Red River
fault zone
and signatures of detached lithosphere
Tengchong is likely caused by a slab
tear of the eastward subducted Indian
under the western Yangtze Craton
plate Ruiqing Zhang1 , Yan Wu1, Zhanyong Gao2, Yuanyuan V. Fu3 , Lian Sun1, Qingju Wu1, and
A detached lithosphere at 660 km
depth below western Yangtze Craton Zhifeng Ding1
1
Institute of Geophysics, China Earthquake Administration, Beijing, China, 2Research Center of Exploration Geophysics, China
Earthquake Administration, Henan, China, 3Institute of Earthquake Science, China Earthquake Administration, Beijing, China
Correspondence to:
R. Zhang,
zrq@cea-igp.ac.cn Abstract We present new constraints on the upper mantle transition zone structure beneath eastern and
southeastern Tibet based on P wave receiver functions for a large broadband data set from two very dense
Citation: seismic arrays. A clear depression of both the 410 km and 660 km discontinuities is detected west of the
Zhang, R., Y. Wu, Z. Gao, Y. V. Fu, L. Sun, Red River fault relative to the east. The correlated topographic variations across the Red River fault are
Q. Wu, and Z. Ding (2017), Upper mantle indicative of temperature changes in the upper mantle above the transition zone, which suggests that the
discontinuity structure beneath eastern
and southeastern Tibet: New constraints fault is a deep-rooted structure that penetrates into the upper mantle and separates Indochina from South
on the Tengchong intraplate volcano China. West of the Red River fault, the transition zone thickness under the Tengchong volcano is found to
and signatures of detached lithosphere be normal compared to the global average. This strongly suggests that the intraplate volcano may originate
under the western Yangtze Craton,
J. Geophys. Res. Solid Earth, 122, from slab tearing of the eastward subducting Indian plate at shallow depths in the upper mantle rather than
doi:10.1002/2016JB013551. from dehydration of a attened plate within the transition zone. Our results further show that the 660 km
discontinuity is signicantly depressed under the western Yangtze Craton and that the transition zone
Received 14 SEP 2016 therefore thickens by up to 20 km. This thickening is suggestive of lowered temperatures associated with a
Accepted 3 FEB 2017
Accepted article online 5 FEB 2017 remnant of detached lithosphere in response to overlying asthenospheric escape ow in and around the
western Yangtze Craton. In addition, we nd that the transition zone thickness beneath much of the Sichuan
Basin is similar to the global average.

1. Introduction
The high, uniform elevation of the Tibetan Plateau is the result of crustal thickening caused by the northward
subduction of Indian lithosphere under the Eurasian plate that began ~50 Myr ago. In this process, eastward
subduction of the Indian plate beneath the Burmese arc has also been conrmed [e.g., Ni et al., 1989]. In contrast
to the at central plateau, the eastern and southeastern (SE) margins bordering the plateau are characterized
both by well-dened steep margins (such as along the Longmenshan fault) and by broad, irregular, and
topographically gradational boundaries [e.g., Burchel et al., 1995; Clark and Royden, 2000]. Evidence of late
Cenozoic crustal shortening is largely absent in eastern and SE Tibet [e.g., Royden et al., 1997].
Tectonic models of surface uplift in eastern and SE Tibet can be roughly cast into two end-members: (1) ductile
ow within the mid-lower crust evacuated from beneath the central plateau and the Sichuan Basin as a barrier
to crustal ow [e.g., Clark and Royden, 2000] and (2) the rotation of crustal blocks in eastern Tibet and south-
eastward extrusion of Indochina relative to South China along the strike-slip Red River fault [e.g., Tapponnier
et al., 1982]. Therefore, the Red River fault is regarded as a potential lithospheric structure, in which case
distinct properties would be detected across the fault [e.g., Schrer et al., 1994; Leloup et al., 1995].
A variety of approaches has been employed to investigate crustal and upper mantle structure below
eastern and SE Tibet. A single model seems difcult to explain crustal deformation of this complex region
[e.g., Yao et al., 2008]. Some studies have proposed that crustal ow and rigid block motion are not irreconcil-
able [e.g., Liu et al., 2014] and that a combination of these two models may play a key role in accommodating
lithospheric deformation in the uplift and eastward expansion of eastern Tibet [Bao et al., 2015]. Eastern and SE
2017. American Geophysical Union.
Tibet have also been suggested as a candidate channel for the ow of asthenosphere material [e.g., Liu et al.,
All Rights Reserved. 2004; Yang et al., 2014; Huang et al., 2015]. Such asthenospheric ow would imply removal of the overlying

ZHANG ET AL. MANTLE TZ IMAGE OF EASTERN AND SE TIBET 1


Journal of Geophysical Research: Solid Earth 10.1002/2016JB013551

mantle lithosphere [Zhang et al., 2010; Hu et al., 2011]. In this scenario, present-day lithospheric tectonics could
reach much greater depths, even down to the mantle transition zone (TZ) below eastern and SE Tibet. So far, no
direct evidence for a removed lithosphere has been reported for this region.
There are few constraints on mantle processes deep beneath eastern and SE Tibet. At both regional and local
scales, travel time P wave tomography has detected strong anomalies beneath parts of eastern and SE Tibet,
but the geometries of these anomalies, particularly beneath the Tengchong volcano, are ambiguous with
considerable variation among models [e.g., Huang and Zhao, 2006; Li et al., 2008; Lei et al., 2009; Huang
et al., 2015]. Because existing tomographic images of these anomalies are inconclusive, the origin of the
Tengchong volcano remains debated [e.g., Wang and Huangfu, 2004; Lei et al., 2009]. In fact, it is hard to
obtain high-resolution images of structures deep beneath eastern and SE Tibet [e.g., Wei et al., 2012;
Huang et al., 2015]. Few seismic stations are available in the Indo-Burma ranges, and most stations are located
east of the Tengchong volcano; therefore, teleseismic rays do not cross well, which leads to the relatively
low-resolution imaging.
In addition to seismic tomography, study of the 410 km and 660 km discontinuities offers another method to
investigate upper mantle structure by measuring its thermal perturbations. Mineral physics experiments have
demonstrated that that the 410 km and 660 km discontinuities arise from phase changes from olivine to
phase and from -spinel to perovskite + magnesiowstite, respectively [e.g., Ito and Takahashi, 1989].
Because the phase transitions of these two discontinuities have opposite Clapeyron slopes (the slopes of
the phase transitions as a function of pressure and temperature), they would be deected toward (away from)
each other at relatively high (low) temperatures. The thickness of the TZ thus correlates with temperature.
Eastern and SE Tibet have been the targets of several receiver function studies. An analysis of receiver func-
tions from a portable small array of ~25 stations revealed variations in TZ thickness beneath the western
Yangtze Craton [Xu et al., 2009]. To the north, Zhang et al. [2010] mapped discontinuity topography across
the Longmenshan thrust fault based on receiver functions from a linear seismic array of 29 stations. To the
south, receiver function images of two discontinuities beneath SE Tibet were obtained from permanent regio-
nal stations in Sichuan and Yunnan provinces, China [Hu et al., 2014]. These small arrays and sparse permanent
stations are insufcient to build more detailed topographic reliefs on the 410 km and 660 km discontinuities,
so the estimated TZ thickness has considerable uncertainty. Regional study of discontinuity topography
across all major geological terranes of eastern and SE Tibet has not previously been conducted.
Here we applied receiver function (RF) analysis techniques to investigate discontinuity topography and the
associated TZ thickness beneath eastern and SE Tibet with unprecedented clarity (Figure 1), which was
enabled by data from two very dense arrays without gaps that were recently deployed in this region.
These data vastly improved coverage over a relatively large geographic area and increased the reliability
of estimated TZ structural variations relative to results based on previously undersampled piercing points
at discontinuity depths with small arrays and sparse stations. We not only provide new constraints on the
mechanism of the Tengchong volcano but also shed light into the depth extent of the Red River fault and
the possible presence of detached lithosphere below the western Yangtze Craton, which are key to under-
standing the tectonic evolution of eastern and SE Tibet.

2. Data and Analysis


Seismic data were obtained from two temporary seismic arrays across eastern and SE Tibet (Figure 1c). The
northern array, installed during 2007 to 2009, consisted of 288 broadband stations spaced at 1030 km inter-
vals, mainly across the Qiangtang and Songpan-Ganzi blocks and the Sichuan Basin [Liu et al., 2014]. The
southern array consisted of 350 broadband stations with an average spacing of ~35 km and was deployed
mainly in SE Tibet by the ChinArray project from 2011 to 2014. Each station of these two arrays was equipped
with a CMG-3ESPC or CMG-40 seismometer and a REFTEK-130 recorder.
To apply the RF technique, we collected events with body wave magnitudes >5.0 that occurred at epicentral
distances of 3090 during the recording periods of the instruments. To ensure good signal-to-noise ratios,
we selected events with clear P wave arrivals and a total of about 1360 earthquakes met our criteria. The
two horizontal components of seismograms were rotated to the great circle path. We applied the time
domain interactive deconvolution method to generate RFs [Ligorria and Ammon, 1999]. We then visually

ZHANG ET AL. MANTLE TZ IMAGE OF EASTERN AND SE TIBET 2


Journal of Geophysical Research: Solid Earth 10.1002/2016JB013551

Figure 1. (a) Map showing topography, major faults, and tectonic units in and around the Tibetan Plateau. (b) The bold black lines depict the major faults that
separate different blocks. (I) Songpan-Ganze Fold Belt, (II) Qiangtang Block, (III) Lhasa Block, and (IV) Indochina Block. Abbreviations are as follows: NJF, Nujiang
fault; JSJF, Jinshajiang fault; RRF, Red River fault; LMSF, Longmenshan fault; XJF, Xiaojiang fault. The yellow triangles show volcanoes in the area; Tengchong volcano
is labeled TCV. The solid circles denote strong earthquakes since 1960 with magnitudes of M > 6.0 (red: M > 7.0; purple: 6.0 < M < 7.0). (c) Location of the two
temporary seismic arrays. The circles and triangles represent the 288 and 350 broadband stations of the two seismic arrays, respectively. The southern seismic array
(triangles) is the ChinArray.

examined individual RFs to remove traces without coherent converted phase arrivals from the Moho. This
selection method produced a nal data set of 195,000 high-quality RFs.
We employed the common-conversion-point (CCP) stacking approach to map topographic variations of the
410 km and 660 km discontinuities and to determine lateral variation in TZ thickness beneath the two arrays
[e.g., Dueker and Sheehan, 1997]. Before stacking, the study region was divided into a three-dimensional pixel
parametric geometry with CCP bins spaced at 15 15 km intervals horizontally and at 1 km depth intervals
from the subsurface to 900 km depth. For each assumed conversion depth, the piercing points and travel
times of the converted phase Ps were calculated by ray tracing with an appropriate model. In this study,
we rst used the 1-D IASP91 model [Kennett and Engdahl, 1991] as the reference model to map time domain
RF signals into discontinuity topography. To account for upper mantle heterogeneity, we also adopted
another model extracted from the 3-D S wave models for East Asia from Rayleigh wave tomography
[Li et al., 2013] by assuming IASP91 Vp/Vs ratios. For simplicity, the shear velocity taken from the 3-D surface
tomography model at each station was specied as the 1-D velocity prole. We stacked the RF amplitudes
that shared the geographic locations of conversion points. Discontinuity topography was also smoothed
within a radius of 3 pixel points, which resulted in an effective horizontal bin size spacing of ~90 km, i.e.,
the same order of the rst Fresnel radius at 410 km (~90 km for 0.23 Hz receiver function). Figure 2 shows

ZHANG ET AL. MANTLE TZ IMAGE OF EASTERN AND SE TIBET 3


Journal of Geophysical Research: Solid Earth 10.1002/2016JB013551

the geographic layout of the piercing points of the Ps conversions from the two discontinuities as well as hit
counts in each bin. The two combined seismic arrays without gaps provide a denser sampling of the TZ under
eastern and SE Tibet than was previously available.

3. Results
3.1. CCP Images
In our CCP images obtained from the IASP91 model, Ps conversions from the two upper mantle discontinu-
ities can easily be identied in most cross sections, as shown in Figure 3. The southernmost cross sections
(Figures 3i and 3j) demonstrate apparent correlation between the 410 km and 660 km discontinuities, which
are both depressed by ~20 km west of the Red River fault relative to east of the fault. To the north
(Figures 3f3h), the depression of the 660 km discontinuity abruptly becomes more pronounced over a lateral
distance of ~200 km between the Red River and Xiaojiang faults, whereas the overlying 410 km discontinuity
returns to its global average depth and is slightly uplifted by a few kilometers particularly at latitude of 26N
(Figure 3g). The localized depression of the 660 km discontinuity disappears gradually toward the north. In
the northernmost cross sections (Figures 3a and 3b), the 660 km discontinuity becomes quite complex, with
multiple peaks at the western end of the sections, which differs from the simple and continuous image to
the east. The strong signal of the multiples is the deeper one which displays substantial depression by ~20
30 km (Figures 3a and 3b).
The depths to the two discontinuities were picked from the CCP images as the average positions of 10 max-
imum positive peaks within the respective depths ranges of 390440 km and 640700 km. Based on the
IASP91 model, we found average depths of 412 km and 663 km for the 410 km and 660 km discontinuities,
respectively, across our study area (Figures 4a and 4c). Variations in discontinuity depth with the IASP91
model are clearly visible in the resulting topographic maps (Figures 5a and 5b). One striking feature of the
topographic maps is coincident depression of the two discontinuities observed in the southwest compared
with the southeast of our study region, as suggested in their CCP proles (Figures 3i and 3j). The coincident
topographic depressions terminate roughly against the trace of the Red River fault at the surface.
Furthermore, a strong localized depression of the 660 km discontinuity appears under the western part of
the Yangtze Craton bounded by the Red River and Xiaojiang faults, while the topographic relief on the
410 km discontinuity does not vary signicantly in that region. Similar topographic variations also appear
at the northwestern edge of our study region. In addition, no signicant topographic anomalies were
detected beneath much of the Sichuan Basin.
The different models used in migration can affect the absolute depths of the 410 km and 660 km discontinu-
ities shown in the CCP image. The mean depths of the two discontinuities obtained from the surface-wave
tomography model are 408 km and 667 km with standard derivations of 6.3 km and 7.8 km, respectively
(Figures 4b and 4d). For most study regions, discontinuity topography from the surface-wave tomography
model is generally similar to that from the IASP91 model (Figures 5d and 5e). We further note that there is
less topographic relief on the 410 km discontinuity west of the Red River fault migrated with the surface-wave
tomography model than with the IASP91 model.
Potential systematic uncertainties introduced by migration with an inaccurate model of crustal and mantle velo-
city structure can be removed by subtracting the depths of the 410 km from the 660 km discontinuities in over-
lapping CCP bins to derive the TZ thickness. We nd similar patterns of TZ thickness by using both the IASP91
model and the surface-wave tomography model (Figures 4c and 4f). The most prominent anomaly in TZ thick-
ness lies beneath the western Yangtze Craton (Figure 4c). It follows a N-S trend over a distance ~400 km and is
approximately 20 km thicker than the regional background and the global average, which corresponds to the
locally depressed 660 km discontinuity shown in Figures 3f3h. Thickening (~20 km) of the TZ also occurs at
the northwestern edges of our study region. In contrast, the TZ thickness west of the Red River fault appears
to be rather typical, similar to the global average of 250 km, as a result of the similar coincident depression of
the two discontinuities (Figures 3i and 3j). A uniform TZ is also imaged beneath much of the Sichuan Basin.

3.2. Comparisons With Previous Studies


Our observed thickest TZ beneath the western Yangtze Craton agrees with the ndings of certain previous RF
studies. Based on data from the MIT-China (YA) seismic array (25 temporary broadband seismic stations)

ZHANG ET AL. MANTLE TZ IMAGE OF EASTERN AND SE TIBET 4


Journal of Geophysical Research: Solid Earth 10.1002/2016JB013551

Figure 2. Distribution of ray piercing points at depths of 410 km and 660 km with the hit counts in each bin. Hit counts of piercing points (black points) (a) at 410 km
depth and (b) at 660 km depth. Each bin is horizontally spaced every 30 km and is 90 90 km wide. The receiver function images along 10 proles (red dotted lines)
are shown in right plot. The bold black lines represent major faults of different blocks. The red triangles show volcanoes in the area.

deployed in SE Tibet from 2003 to 2005, the receiver function study of Xu et al. [2009] indicated that the TZ
thickens by up to 20 km beneath several stations near the Red River fault compared with the global average.
However, our location for the thickest TZ is slightly offset from the anomalous TZ reported by Xu et al. [2009].
This spatial offset is likely due to the stacking technique used by Xu et al. [2009], in which RFs were not binned
based on the geographic piercing points of the Ps phases because of the insufcient number of stations. The
anomalously thick TZ inferred by Xu et al. [2009] would be shifted toward the southeast because a large pro-
portion of incoming events were sourced from the southwestern Pacic Ocean. Recently, using some perma-
nent regional network stations, P wave RFs mapped a depressed 660 km discontinuity under the western
Yangtze Craton [Hu et al., 2014], in agreement with our observed topographic anomaly in that area.
We further note that there are disagreements between our results and those of some RF studies. With a small
seismic array, Zhang et al. [2010] found a depressed 410 km discontinuity toward the west across the
Longmenshan fault, and associated thinner-than-normal TZ below eastern Tibet, in contrast to the thicker-
than-normal TZ below the Sichuan Basin. The locally depressed 410 km discontinuity along the
Longmenshan fault was also observed by Hu et al. [2014]. However, we did not detect an obvious depression
of the 410 km discontinuity with a thinner TZ along the western margin of the Sichuan Basin. In comparison,
our CCP images have a greatly improved sampling of the Sichuan Basin and adjacent regions; therefore, the
differences between our discontinuity topography and those from previous RF studies are likely the result of
this higher sampling density.
In addition, we note that the strong depression of the 660 km discontinuity bounded by the Red River and
Xiaojiang faults, as inferred from the depth section of the CCP stacked RFs (Figure 6b), correlates with the
high-velocity structure in the TZ (Figure 6a) revealed by a recent P wave tomography study by Huang et al.
[2015]. However, the localized increase in TZ thickness across our study region seems contradictory with a
recent regional tomography study that indicated a continuous high-velocity anomaly across a broad area
from the Burmese arc northward to the Kunlun fault and eastward to the Xiaojiang fault [Lei and Zhao, 2016].

4. Discussion
4.1. The Red River Fault: A Deep-Rooted Lithospheric Structure
The Red River fault is one of the most prominent strike-slip faults in Southeast Asia. It extends over 900 km
from the eastern Himalayan syntaxis to the Tonkin Gulf, and as far as the South China Sea. As mentioned
in the introduction, the NW-SE striking Red River fault is considered a major tectonic structure along which

ZHANG ET AL. MANTLE TZ IMAGE OF EASTERN AND SE TIBET 5


Journal of Geophysical Research: Solid Earth 10.1002/2016JB013551

Figure 3

ZHANG ET AL. MANTLE TZ IMAGE OF EASTERN AND SE TIBET 6


Journal of Geophysical Research: Solid Earth 10.1002/2016JB013551

Indochina was extruded toward the SE as a result of the India-Asia collision [e.g., Tapponnier et al., 1982;
Schrer et al., 1994]. Different patterns of velocity anomalies have been imaged across and along the
Red River fault with seismic tomography. Some P and S wave tomography from local earthquakes show
the Red River fault as a boundary between low- and high-velocity anomalies at around 4050 km depth
[Huang et al., 2002; Wang et al., 2003]. Recent inversion of Rayleigh wave dispersion and RFs have
produced evidence of abrupt changes in shear-wave velocity at a depth of 5080 km across the Red
River fault [Bao et al., 2015; Zheng et al., 2016]. These ndings suggest that the shear zone cuts through
the entire crust and may extend into the upper mantle. Instead of contrasting velocity anomalies across
the Red River fault, other tomography studies based on local and teleseismic data have shown a
prominent P wave low-velocity anomaly along the fault zone from the lower crust to ~100 km depth
[Lei et al., 2009]. Similar low-velocity anomalies at the top of the upper mantle have also been revealed
beneath the southeastern half of the Red River fault in northern Vietnam from local tomography using P
and Pn waves [Huang et al., 2013].
In this study, we found a distinct contrast in topographic variations of upper mantle discontinuities across the
Red River fault. Both the 410 km and 660 km discontinuities are depressed by 1020 km west of the Red River
fault with respect to the east. The correlated depression of the two discontinuities may be caused by an
unmodeled low-velocity anomaly in the upper mantle. An average depression of ~15 km in the 410 km dis-
continuity can be explained either by a strong (48%) localized low-velocity anomaly at the lithosphere-
asthenosphere boundary or by an ~1.5% anomaly across the entire upper mantle above 410 km depth
[Wang and Niu, 2011]. At present, we cannot yet constrain the exact depth extent of the low-velocity anomaly
west of the Red River fault, but both cases indicate that this fault, as a deep-rooted lithospheric boundary,
penetrates at least to the upper mantle at the boundary between Indochina and South China.

4.2. Origin of the Tengchong Volcano


In spite of topographic changes in the two discontinuities across the Red River fault, the TZ thickness does not
vary signicantly between the Indochina and South China blocks, except for the western Yangtze Craton. In the
mantle, lateral variation in temperature would affect TZ thickness. The expected magnitude of such a thermal
effect is about 70100C per 10 km change in thickness of the TZ [e.g., Helffrich, 2000]. The fact that the TZ has
normal thickness implies that the associated thermal anomaly must be less than 70100C, which is presumed
to be the limit of its temperature stability [Zhao et al., 2010]. We therefore conclude that there is no subducting
slab present in the TZ below most of SE Tibet, except for the western Yangtze Craton.
The TZ structure west of the Red River fault has important implications for the mechanism of the Tengchong
volcano (25N, 98E), which is located on the northeastern margin of the India-Asia collision zone and is about
450 km from the eastern boundary of the Indian plate. This active volcano has shown multiple phases of
activity from the Miocene to the Quaternary [Jiang, 1998]. Many studies of the Tengchong volcano and adja-
cent regions have been conducted, but the origin of this volcanism is not yet well resolved. Seismicity
beneath the Indo-Burma ranges indicates that an eastward dipping Wadati-Benioff zone extends to a maxi-
mum depth of about 180 km, but it does not extend as far east as the Tengchong volcano [e.g., Ni et al., 1989].
Consistent with the seismicity result, regional P wave tomography reveals a linear high-velocity anomaly to a
depth of about 200 km under the Indo-Burma ranges [Huang and Zhao, 2006; Li et al., 2008], which conrms
the presence of eastward-subducted Indian lithosphere [Li et al., 2008]. To the east, the Tengchong volcano is
associated with a more complex distribution of velocity anomalies detected in tomography studies. Some
regional tomography studies using teleseismic data sets indicate high-velocity anomalies in the TZ centered
~400 km east of the Tengchong volcano and are isolated from the anomalous high-velocity structure imaged
at a depth of 200 km beneath the Indo-Burma ranges [Huang and Zhao, 2006; Li et al., 2008; Li and Van der
Hilst, 2010], whereas the uppermost ~150 km of the mantle beneath the volcano shows anomalously low
velocity [Li et al., 2008]. In contrast, other tomography studies both local and teleseismic data sets reveal a
pronounced low-velocity volume extending from the subsurface to 300400 km depth directly above a
high-velocity anomaly in the TZ just beneath the Tengchong volcano [Lei et al., 2009; Yang et al., 2014;

Figure 3. CCP stacked images based on the IASP91 model across latitudes from 23N to 32N with a longitudinal origin of 103E (corresponding proles shown in
Figure 2b). Ellipses indicate local abrupt depressions of the 660 km discontinuity observed in the middle and northern parts of our study region. The red triangle
indicates the Tengchong Volcano (TCV). The black arrows denote the Red River fault (RRF), Xiaojiang fault (XJF), and Longmenshan fault (LMSF). The inverted triangles
indicate the seismic stations along each prole.

ZHANG ET AL. MANTLE TZ IMAGE OF EASTERN AND SE TIBET 7


Journal of Geophysical Research: Solid Earth 10.1002/2016JB013551

Figure 4. Distribution of 410 km discontinuity depth picks between 390 and 430 km based on (a) the IASP91 model and
(b) on the surface-wave tomography model. Distribution of 660 km discontinuity depth picks between 640 and 700 km
based on (c) the IASP91 model and (d) on the surface-wave tomography model. The best t normal distribution is indicated
as a black line.

Huang et al., 2015]. The high-velocity anomaly in the TZ is continuously connected with the eastward high-
velocity structure in the upper mantle below the Burmese arc [Huang et al., 2015]. The former ndings sug-
gest that the origin of the Tengchong volcano is not deep [e.g., Wang et al., 2003], whereas the latter results
indicate that the volcano originates from dehydration of the Indian plate in the TZ and upwelling of hydrated
mantle from the attened plate [e.g., Lei et al., 2009]. The latter model has been invoked to explain the origin
of the Changbaishan intraplate volcanism in NE China. It should be noted that the deep seismicity clusters
beneath the Changbaishan volcanism reach a maximum depth of ~600 km compared with the shallow depth
of ~180 km under the Tengchong volcano.
Our inference of a uniform TZ west of the Red River fault, as an indicator of the absence of a subducting slab,
would require that the eastward subducting Indian plate does not continuously extend far beyond the
Tengchong volcano. This interpretation is in agreement with a previous global P wave tomographic study
by Replumaz et al. [2010]. In that study, a high-velocity anomaly is imaged with a relatively shallow angle
to around 200 km depth, where it steepens to a depth of 400 km beneath the Burma arc, and is mainly con-
ned west of 95E between the Indo-Burma suture and the Sagaing fault. Combined with this tomographic
model, the absence of a thermal anomaly in the TZ beneath the Tengchong volcano leads us to speculate
that slab tearing of the subducted Indian plate occurs at shallower depths above 410 km. This slab tearing
may be the trigger of localized upwelling that would feed the Tengchong volcano (Figure 7b). An analysis
of geochemical and geochronological data and model simulations also suggest that the Tengchong volcanic
eld magmas may originate from partial melting of the enriched asthenospheric mantle wedge associated
with slab detachment of Indian continental lithosphere [Guo et al., 2015]. We suggest that this slab tearing
is most likely to occur at a depth of ~200 km, where the deepest seismicity of the Wadati-Benioff zone along
the Burma arc terminates. Slab tearing is expected to occur at oceanic slab-slab junctions [e.g., Obayashi et al.,
2009; Zhang et al., 2016]. The mechanism of a continental slab tearing, however, requires further study.

4.3. The Western Yangtze Craton: Detached Lithosphere in the TZ


In our study region, the thickest TZ occurs below the western Yangtze Craton where the 660 km discontinuity
is signicantly depressed by ~20 km. For the phase change across the 660 km discontinuity, the measured

ZHANG ET AL. MANTLE TZ IMAGE OF EASTERN AND SE TIBET 8


Journal of Geophysical Research: Solid Earth 10.1002/2016JB013551

Figure 5. Discontinuity topography and transition zone thickness maps: (a) 410 km discontinuity depth, (b) 660 km discontinuity depth, and (c) thickness of transition
zone (TZ) migrated based on the IASP91 model for the study area, (d) 410 km discontinuity depth, (e) 660 km discontinuity depth, and (f) thickness of transition
zone (TZ) based on the surface-wave tomography model used to correct for mantle heterogeneity [Li et al., 2013]. The purples circles denote earthquakes that have
occurred since 1960 within 90100E and 1832N with focal depths of more than 50 km. Other labeling is the same as that of Figure 2.

ZHANG ET AL. MANTLE TZ IMAGE OF EASTERN AND SE TIBET 9


Journal of Geophysical Research: Solid Earth 10.1002/2016JB013551

Figure 6. (a) Comparisons between a tomographic image of P wave velocity variation along the latitude of 25.5N from
Huang et al. [2015] and a depth section of the CCP stacked receiver functions migrated with the surface-wave tomography
model [Li et al., 2013]) along the same latitude. (b) At depths between 350 and 510 km and between 511 and 800 km, all
receiver functions are stacked based on their piercing points that fall within the same bin at 410 km and 660 km, respectively.
The circles denote earthquakes that have occurred since 1960 between 24N and 26N with magnitudes of M > 5.0. (c) Depths
of the two discontinuities picked from the CCP stacks. The depths to the 410 km and 660 km discontinuities are indicated
on the left and right, respectively. The gray bars show 1 error bounds for the mean depths (circles) of the two discontinuities
estimated from bootstrapping. The maximum uncertainty of discontinuity depths typically occurs at the two edges of the
prole where the image may be contaminated due to insufcient data.

Clapeyron slope varies from 2.8 MPa/K to 1.3 MPa/K [e.g., Ito and Takahashi, 1989; Fei et al., 2004]. Using a
Clapeyron slope of 2.0 MPa/K, we calculate a thermal anomaly of about 320 K for 20 km of depression. This
result matches those of a regional P wave tomography study of Southeast Asia, which have indicated a high-
velocity anomaly of ~1.5% in the TZ centered east of the Red River fault [Li and Van der Hilst, 2010]. Given a
temperature sensitivity of 0.43%/100 K [Cammarano et al., 2003], an ~1.5% anomaly requires a thermal per-
turbation of about 350 K, which is consistent with expectations based on our RF results.
High-velocity anomalies in the TZ are usually interpreted in terms of low temperatures associated with a sub-
ducted plate or removed lithosphere that sank from above. The presence of a high-velocity structure in the
TZ east of the Red River fault has been interpreted as related to subduction along a different arc, such as the

ZHANG ET AL. MANTLE TZ IMAGE OF EASTERN AND SE TIBET 10


Journal of Geophysical Research: Solid Earth 10.1002/2016JB013551

Figure 7. Comparisons between models for the origin of the Tengchong volcano (TCV). (a) A schematic cross section
showing the model proposed based on tomographic image of P wave velocity variation by Lei et al. [2009]. The circles
denote earthquakes that have occurred since 1960 between 24N and 26N with magnitudes of M > 4.0. The triangle
indicates the Tengchong intraplate volcano (TCV). For reference, we also show key geological units and boundaries.
Abbreviations are as follows: NJF, Nujiang fault; RRF, Red River fault; XJF, Xiaojiang fault. (b) A schematic cross section
showing our proposed model. The upwelling shown in red is associated with tearing of the Indian slab. The two dashed
lines in each panel represent the 410 km and 660 km discontinuities. The subducting Indian slab is shown in brown. There is
no vertical exaggeration in these two cross sections except for surface topography in the top plot.

Late Mesozoic South China Trench over approximately 85 Ma [Li and Van der Hilst, 2010], based on recon-
struction of the formation of arcs and back-arc basins in Southeast Asia [Honza and Fujioka, 2004]. It has been
estimated that in Southeast Asia, the sinking of cold lithosphere with little lateral advection in a mantle refer-
ence frame takes place at rates of ~2 cm/yr below 700 km and ~5 cm/yr above 700 km [Replumaz et al., 2004].
Assuming a sinking rate of ~3 cm/yr, the slab geometry between 600 and 900 km depth corresponds to initia-
tion of plate subduction at around 1530 Ma, which is inconsistent with the estimate of 85 Ma suggested by Li
and Van der Hilst [2010].
Alternatively, the thick TZ anomaly beneath the western Yangtze Craton may imply a remnant of detached
mantle lithosphere. The Yangtze Craton is a very stable Precambrian craton, separated from the North
China Craton by the Qingling-Dabie orogenic belt to the north and from the Tibetan Plateau by the
Longmenshan and Red River faults to the west and southwest. S wave RF results indicate that the lithospheric
thickness under the western Yangtze Craton is about ~80 km, whereas it increases northward to 140180 km
under the Sichuan Basin and southeastward to 140160 km [Hu et al., 2011]. In contrast to the high-velocity
anomaly that covers most of the Yangtze Craton and extends as deep as 300 km, regional tomographic
images suggest the presence of a low-velocity anomaly at 60250 km depth around the Xiaojiang fault
[Huang et al., 2015]. The high-velocity volume is interpreted as a root of the stable Yangtze Craton.
Although there are differences in estimations of the lithospheric thickness of the Yangtze Craton between

ZHANG ET AL. MANTLE TZ IMAGE OF EASTERN AND SE TIBET 11


Journal of Geophysical Research: Solid Earth 10.1002/2016JB013551

the results of S wave RFs and regional tomography, both studies have indicated a lateral decrease in litho-
spheric thickness from the Sichuan Basin toward the southwest. This lateral decrease strongly suggests a rela-
tively thin lithosphere beneath the western Yangtze Craton. Therefore, it is possible that the base of the
lithosphere under the western Yangtze Craton has been removed. We suggest that the TZ under the western
Yangtze Craton is a resting ground for the detached lithosphere (Figure 7b) and that this process is what led
to its increased thickness rather than slab subduction ~85 Ma [Li and Van der Hilst, 2010]. Such detachment
may be facilitated by mantle ow extruded from the Tibetan Plateau. The extensive low-velocity anomalies
imaged at 60250 km depth around the Xiaojiang fault may be indicative of channels in the upper mantle
that allow southeastward extrusion of asthenospheric material under the Tibetan Plateau and maintain the
balance of the Plateau [Huang et al., 2015].

4.4. Complex 660 km Discontinuity


The thicker TZ is also detected at the northwestern edge of the study region, where multiple conversions and
an ~2030 km depression of the 660 km discontinuity are present, as described above (Figures 3a and b).
Multiple conversions originating from the 660 km discontinuity have been reported from many studies
[e.g., Niu and Kawakatsu, 1998; Chen and Ai, 2009]. Nonolivine minerals such as garnet may play an impor-
tant role at depths around 660 km [Vacher et al., 1998]. However, because multiple conversions are
observed at the western edge of the proles (Figures 3a and 3b), where the image may be contaminated
because of insufcient data, such multiples should be interpreted with caution [Chen and Ai, 2009]. The multi-
ples are not observed beneath the Western Yangtze Craton, where the depression is also prominent.
Therefore, our preferred interpretation of the 660 km discontinuity in the northwestern part of the study area
is as a single discontinuity.
Similar to the proposed mechanism for the TZ thickness anomaly under the western Yangtze Craton, the
thicker TZ in the northwestern part of the study area near the eastern Himalayas also indicates the presence
of detached lithosphere that has sunk from above. To the west of our study region, the RF results reported by
Singh and Kumar [2009] reveal a thick TZ beneath much of the eastern Himalayas and southern Tibet from
portable array data. They conclude that the thickening of the TZ is mainly due to the presence of removed
lithosphere stagnating in the TZ, which is invoked either by convection or gravity removal of the thickened
lithosphere associated with the continuous subduction of the Indian plate beneath the Asian plate. The loca-
lized anomaly in TZ thickness we have detected near the northwestern edge of our study region requires
further investigation in order to understand its nature.

4.5. TZ Structure Beneath the Sichuan Basin


In addition, we found that the Sichuan Basin is characterized by a normal TZ, with thickness close to the esti-
mated global average. S wave tomography results have indicated a high-velocity anomaly within the TZ
below the Sichuan Basin [Lebedev and Nolet, 2003; Obrebski et al., 2012]. This high-velocity anomaly has been
interpreted as a lithospheric remnant of the South China block-North China Craton collision [Lebedev and
Nolet, 2003], and it has been suggested that it plays a role in the stability of the Sichuan Basin [Obrebski
et al., 2012]. In contrast, recent regional tomography has revealed extensive P wave low-velocity anomalies
in both the TZ and the lower mantle below the Sichuan Basin, which is supposed to be driven by subduction
[Huang et al., 2015]. Inconsistent with these previous studies, we found no indication of a thermal anomaly
within the TZ beneath much of the Sichuan Basin.

5. Conclusions
We investigated the TZ structure beneath eastern and SE Tibet with unprecedented detail by using RF data
recorded by two very dense seismic arrays. Our CCP stacked imaging reveals a correlated depression of both
the 410 and 660 km discontinuities west of the Red River fault compared with the east. The variations in dis-
continuity topography across the Red River fault indicate that the average temperature of the upper mantle
above the TZ depth is hotter in the west than in the east, which suggests that the fault zone is a deep-rooted
structure that penetrates into the upper mantle and separates Indochina from South China. We found that
the TZ thickness does not vary signicantly across our study region except for beneath the western
Yangtze Craton and at the northwestern edge. Lack of evidences for anomalous TZ thickness beneath the
Tengchong volcano suggests that this intraplate volcano is unlikely to originate from dehydration of a

ZHANG ET AL. MANTLE TZ IMAGE OF EASTERN AND SE TIBET 12


Journal of Geophysical Research: Solid Earth 10.1002/2016JB013551

attened slab in the TZ but may originate from slab tearing of the eastward subducting Indian plate at shal-
low depths in the upper mantle. The thickest TZ of the western Yangtze Craton may represent a remnant of
detached lithosphere resting within the TZ, facilitated by overlying asthenospheric escape ow in that region.
In contrast, there is no evidence of a thermal anomaly within the TZ beneath much of the Sichuan Basin.

Acknowledgments References
The waveform data were provided by
the China Seismic Array Data Bao, X., X. Sun, M. Xu, D. W. Eaton, X. Song, L. Wang, Z. Ding, N. Mi, H. Li, and D. Yu (2015), Two crustal low-velocity channels beneath SE Tibet
Management Center at the Institute of revealed by joint inversion of Rayleigh wave dispersion and receiver functions, Earth Planet. Sci. Lett., 415, 1624.
Geophysics, China Earthquake Burchel, B., Z. Chen, Y. Liu, and L. Royden (1995), Tectonics of the Longmen Shan and adjacent regions, central China, Int. Geol. Rev., 37, 661735.
Administration. We are very grateful to Cammarano, F., S. Goes, P. Vacher, and D. Giardini (2003), Inferring upper mantle temperatures from seismic velocities, Phys. Earth Planet. Int.,
Barbara Romanowicz and an anon- 138, 197222.
ymous reviewer whose suggestions Chen, L., and Y. Ai (2009), Discontinuity structure of the mantle transition zone beneath the North China Craton from receiver function
signicantly improved the manuscript. migration, J. Geophys. Res., 114, B06307, doi:10.1029/2008JB006221.
We would also like to thank L.P. Zhu for Clark, M. K., and L. H. Royden (2000), Topographic ooze: Building the eastern margin of Tibet by lower crustal ow, Geology, 28, 703706.
providing the deconvolution code. This Dueker, K., and A. Sheehan (1997), Mantle discontinuity structure from midpoint stacks of converted P and S waves across the Yellowstone
research was supported by the National hotspot track, J. Geophys. Res., 102(B4), 83138327, doi:10.1029/96JB03857.
Science Foundation of China (grant Fei, Y., J. Van Orman, J. Li, W. van Westrenen, C. Sanloup, W. Minarik, K. Hirose, T. Komabayashi, M. Walter, and K. Funakoshi (2004),
41474089) and the China National Experimentally determined postspinel transformation boundary in Mg2SiO4 using MgO as an internal pressure standard and its geo-
Special Fund for Earthquake Scientic physical implications, J. Geophys. Res., 109, B02305, doi:10.1029/2003JB002562.
Research in Public Interest (201008001 Guo, Z., Z. Cheng, M. Zhang, L. Zhang, X. Li, and J. Liu (2015), Post-collisional high-K calc-alkaline volcanism in Tengchong volcanic eld, SE
and 201308011). Tibet: Constraints on Indian eastward subduction and slab detachment, J. Geol. Soc. London., 172, 5054.
Helffrich, G. (2000), Topography of the transition zone seismic discontinuity, Rev. Geophys., 38, 141158, doi:10.1029/1999RG000060.
Honza, E., and K. Fujioka (2004), Formation of arcs and backarc basins inferred from the tectonic evolution of southeast Asia since the Late
Cretaceous, Tectonophysics, 384(1), 2353, doi:10.1016/j.tecto.2004.02.006.
Hu, J., X. Q. Xu, H. Y. Yang, L. M. Wen, and G. Q. Li (2011), S receiver function analysis of the crustal and lithospheric structures beneath eastern
Tibet, Earth Planet. Sci. Lett., 306(12), 7785.
Hu, J., H. Yang, G. Li, and H. Peng (2014), Seismic upper mantle discontinuities beneath Southeast Tibet and geodynamic implications,
Gondwana Res., 28, 10321047.
Huang, J., and D. Zhao (2006), High-resolution mantle tomography of China and surrounding regions, J. Geophys. Res., 111, B09303,
doi:10.1029/2005JB004066.
Huang, J., D. Zhao, and S. Zheng (2002), Lithospheric structure and its relationship to seismic and volcanic activity in southwest China,
J. Geophys. Res., 107(B10), 2255, doi:10.1029/2000JB000137.
Huang, H. H., Z. J. Xu, Y. M. Wu, X. Song, B. S. Huang, and M. N. Le (2013), First local seismic tomography for Red River shear zone, northern
Vietnam: Stepwise inversion employing crustal P and Pn waves, Tectonophysics, 584, 230239.
Huang, Z., P. Wang, M. Xu, L. Wang, Z. Ding, Y. Wu, M. Xu, N. Mi, D. Yu, and H. Li (2015), Mantle structure and dynamics beneath SE Tibet
revealed by new seismic images, Earth Planet. Sci. Lett., 411, 100111.
Ito, E., and E. Takahashi (1989), Postspinel transformations in the system Mg2SiO4Fe2SiO4 and some geophysical implications, J. Geophys.
Res., 94(B8), 10,63710,646, doi:10.1029/JB094iB08p10637.
Jiang, C. (1998), Distribution characteristics of Tengchong volcanoes in the Cenozoic Era [in Chinese with English abstract], J. Seismol. Res.,
21(4), 309319.
Kennett, B., and E. Engdahl (1991), Travel times for global earthquake location and phase identication, Geophys. J. Int., 105, 429465.
Lebedev, S., and G. Nolet (2003), Upper mantle beneath Southeast Asia from S velocity tomography, J. Geophys. Res. 108(B1), 2048,
doi:10.1029/2000JB000073.
Lei, J., and D. Zhao (2016), Teleseismic P-wave tomography and mantle dynamics beneath Eastern Tibet, Geochem. Geophys. Geosyst., 17,
18611884, doi:10.1002/2016GC006262.
Lei, J., D. Zhao, and Y. Su (2009), Insight into the origin of the Tengchong intraplate volcano and seismotectonics in southwest China from
local and teleseismic data, J. Geophys. Res., 114, B05302, doi:10.1029/2008JB005881.
Leloup, P. H., R. Lacassin, P. Tapponnier, U. Schrer, D. Zhong, X. Liu, L. Zhang, S. Ji, and P. T. Trinh (1995), The Ailao Shan-Red River shear zone
(Yunnan, China), Tertiary transform boundary of Indochina, Tectonophysics, 251, 310.
Li, C., and R. D. Van der Hilst (2010), Structure of the upper mantle and transition zone beneath southeast Asia from travel time tomography,
J. Geophys. Res., 115, B07308, doi:10.1029/2009JB006882.
Li, C., R. D. Van der 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, 157168, doi:10.1016/j.epsl.2008.07.016.
Li, Y., Q. Wu, J. Pan, F. Zhang, and D. Yu (2013), An upper-mantle S-wave velocity model for East Asia from Rayleigh wave tomography, Earth
Planet. Sci. Lett., 377378, 367377.
Ligorria, J., and C. Ammon (1999), Iterative deconvolution and receiver-function estimation, Bull. Seismol. Soc. Am., 89, 13951400.
Liu, M., X. Cui, and F. Liu (2004), Cenozoic rifting and volcanism in eastern China: A mantle dynamic link to the Indo-Asian collision?
Tectonophysics, 393, 2942.
Liu, Q., R. D. Van der Hilst, Y. Li, H. Yao, J. Chen, B. Guo, S. Qi, J. Wang, H. Huang, and S. Li (2014), Eastward expansion of the Tibetan Plateau by
crustal ow and strain partitioning across faults, Nat. Geosci., 7, 361365.
Ni, J. F., M. Guzman-Speziale, M. Bevis, W. E. Holt, T. C. Wallace, and W. Seager (1989), Accretionary tectonics of Burma and the three-
dimensional geometry of the Burma subduction zone, Geology, 17, 6871.
Niu, F., and H. Kawakatsu (1998), Determination of the absolute depths of the mantle transition zone discontinuities beneath China: Effect of
stagnant slabs on mantle transition zone discontinuities, Earth Planets Space, 50, 965975.
Obayashi, M., J. Yoshimitsu, and Y. Fukao (2009), Tearing of stagnant slab, Science, 324, 11731175.
Obrebski, M., R. M. Allen, F. Zhang, J. Pan, Q. Wu, and S. H. Hung (2012), Shear wave tomography of China using joint inversion of body and
surface wave constraints, J. Geophys. Res., 117, B01311, doi:10.1029/2011JB008349.
Replumaz, A., H. Krason, R. D. van der Hilst, J. Besse, and P. Tapponnier (2004), 4-D evolution of SE Asias mantle from geological
reconstructions and seismic tomography, Earth Planet. Sci. Lett., 221, 103115.

ZHANG ET AL. MANTLE TZ IMAGE OF EASTERN AND SE TIBET 13


Journal of Geophysical Research: Solid Earth 10.1002/2016JB013551

Replumaz, A., A. M. Negredo, S. Guillot, and A. Villaseor (2010), Multiple episodes of continental subduction during India/Asia convergence:
Insight from seismic tomography and tectonic reconstruction, Tectonophysics, 483, 125134.
Royden, L. H., B. C. Burchel, R. W. King, E. Wang, Z. Chen, F. Shen, and Y. Liu (1997), Surface deformation and lower crustal ow in eastern
Tibet, Science, 276, 788790.
Schrer, U., L.-S. Zhang, and P. Tapponnier (1994), Duration of strike-slip movements in large shear zones: The Red River belt, China, Earth
Planet. Sci. Lett., 126, 379397.
Singh, A., and M. R. Kumar (2009), Seismic signatures of detached lithospheric fragments in the mantle beneath eastern Himalaya and
southern Tibet, Earth Planet. Sci. Lett., 288, 279290.
Tapponnier, P., G. Peltzer, A. Y. Le Dain, R. Armijo, P. Jussieu, and P. Cobbold (1982), Propagating extrusion tectonics in Asia: New insights
from simple experiments with plasticine, Geology, 10, 611616.
Vacher, P., A. Mocquet, and C. Sotin (1998), Computations of seismic proles from mineral physics: The importance of the non-olivine
components for explaining the 660 km depth discontinuity, Phys. Earth Planet. Inter., 106, 275298.
Wang, C., and G. Huangfu (2004), Crustal structure in Tengchong Volcano-Geothermal Area, western Yunnan, China, Tectonophysics, 380,
6987.
Wang, C., W. W. Chan, and W. D. Mooney (2003), Three-dimensional velocity structure of crust and upper mantle in southwestern China and
its tectonic implications, J. Geophys. Res., 108(B9), 2442, doi:10.1029/2002JB001973.
Wang, X., and F. Niu (2011), Imaging the mantle transition zone beneath eastern and central China with CEArray receiver functions,
Earthquake Sci., 24, 6575.
Wei, W., J. Xu, D. Zhao, and Y. Shi (2012), East Asia mantle tomography: New insight into plate subduction and intraplate volcanism, J. Asian
Earth Sci., 60, 88103.
Xu, Q., J. Zhao, Z. Cui, and M. Liu (2009), Structure of the crust and upper mantle beneath the southeastern Tibetan Plateau by P and S
receiver fnctions [in Chinese with English abstract], Chin. J. Geophys., 52(12), 30013008.
Yang, T., J. Wu, L. Fang, and W. Wang (2014), Complex Structure beneath the Southeastern Tibetan Plateau from teleseismic P-wave
tomography, Bull. Seismol. Soc. Am., 104, 10561069.
Yao, H., C. Beghein, and R. D. Van der Hilst (2008), Surface wave array tomography in SE Tibet from ambient seismic noise and two-station
analysisII. Crustal and upper-mantle structure, Geophys. J. Int., 173, 205219.
Zhang, R., Z. Gao, Q. Wu, Z. Xie, and G. Zhang (2016), Seismic images of the mantle transition zone beneath Northeast China and the
Sino-Korean craton from P-wave receiver functions, Tectonophysics, 675, 159167.
Zhang, Z., X. Yuan, Y. Chen, X. Tian, R. Kind, X. Li, and J. Teng (2010), Seismic signature of the collision between the east Tibetan escape ow
and the Sichuan Basin, Earth Planet. Sci. Lett., 292, 254264.
Zhao, J., et al. (2010), The boundary between the Indian and Asian tectonic plates below Tibet, Proc. Natl. Acad. Sci. U. S. A., 107, 11,22911,233.
Zheng, C., Z. Ding, and X. Song (2016), Joint inversion of surface wave dispersion and receiver function for crustal and uppermost mantle
structure in Southeast Tibetan Plateau [in Chinese with English abstract], Chin. J. Geophys., 59(9), doi:10.6038/cjg20160901.

ZHANG ET AL. MANTLE TZ IMAGE OF EASTERN AND SE TIBET 14

You might also like