2018 GC 007515

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Geochemistry, Geophysics, Geosystems

RESEARCH ARTICLE Ambient Noise Tomography of the Shandong Province and


10.1029/2018GC007515
its Implication for Cenozoic Intraplate Volcanism
Key Points:
• There are significant differences
in Eastern China
between southern and northern Cuilin Li1,2,3 , Chuanxu Chen4, Dongdong Dong1,2,3, Ayodeji Paul Kuponiyi5,6 ,
portions of the Yishu fault zone in the
Shandong province of China Stan E. Dosso6 , and Daolei Su7
• The Yishu fault zone plays an 1
important role in the Cenozoic Key Laboratory of Marine Geology and Environment, Institute of Oceanology, Chinese Academy of Sciences, Qingdao,
continental extension, lithospheric China, 2Laboratory for Marine Geology, Qingdao National Laboratory for Marine Science and Technology, Qingdao, China,
3
thinning, and intraplate volcanism Center for Ocean Mega-Science, Chinese Academy of Sciences, Qingdao, China, 4Institute of Deep-Sea Science and
• Asthenosphere upwelling is focused
Engineering, Chinese Academy of Sciences, Sanya, China, 5Geological Survey of Canada, Pacific Geoscience Centre, Sidney,
into localized features and supplies
the Cenozoic volcanoes of the British Columbia, Canada, 6School of Earth and Ocean Sciences, University of Victoria, Victoria, British Columbia, Canada,
7
eastern China continental margin Earthquake Administration of Jinan City, Jinan, China

Supporting Information: Abstract New images of shear wave velocity (VS) structure of the crust and uppermost mantle in the
• Supporting Information S1
Shandong province of China are presented based on inversion of ambient seismic noise data from 47
broadband stations. Interstation Rayleigh wave phase-velocity dispersion curves for periods of 1–40 s are
Correspondence to:
obtained from noise cross correlations between stations. Two-dimensional phase-velocity maps at periods of
C. Li and C. Chen, 5–30 s are obtained by tomographic inversion of these dispersion curves, which are then inverted for 3-D Vs
cuilinli@qdio.ac.cn; heterogeneity to ~80-km depth. Our results show clear indications of Moho uplift and crustal thinning
chencx@idsse.ac.cn
beneath the northern segment of the Yishu fault zone and the Jiaobei uplift, thus providing seismological
evidence for the steep geometry and deep penetration of the fault system. Instead of large-scale, low-velocity
Citation: anomalies beneath eastern China as imaged by previous regional tomography, our high-resolution Vs results
Li, C., Chen, C., Dong, D., Kuponiyi, A. P.,
Dosso, S. E., & Su, D. (2018). Ambient
reveal fine-scale focused low-velocity anomalies beneath the isolated Cenozoic basalts. There are strong
noise tomography of the Shandong spatial correlations between the narrow low Vs zones in the uppermost mantle and the thinned crust, as well
province and its implication for as the locations of the isolated Cenozoic volcanism in the study area. Our results suggest that the
Cenozoic intraplate volcanism in
eastern China. Geochemistry, Geophysics,
asthenosphere upwelling possibly originating from the stagnant Pacific slab is focused into localized features
Geosystems, 19, 3286–3301. https://doi. at 50- to 80-km depth that supply the isolated Cenozoic volcanoes of the eastern China continental margin.
org/10.1029/2018GC007515

Received 1 MAR 2018


Accepted 16 AUG 2018
Accepted article online 23 AUG 2018 1. Introduction
Published online 19 SEP 2018
The genesis of intraplate volcanism remains a matter of considerable debate (Faccenna et al., 2010). In the
eastern China continental margin, Cenozoic intraplate basalts are widely spread, extending from the
Heilongjiang province in the north to Hainan Island in the south, which represents one of the most presently
active regions of intraplate volcanism in the world (Sakuyama et al., 2013; Tang et al., 2013; Xu, 2014). Most of
the Cenozoic basalts are distributed in extensional basins, rift systems, and their adjacent areas. Mantle
upwelling and decompression melting at shallow depths have previously been invoked to explain
Cenozoic magmatism in eastern China (Huang & Zhao, 2006; Kameyama & Nishioka, 2012; Maruyama
et al., 2009).
The Pacific plate is subducting northwestward beneath the Eurasian plate, and the slab has remained in the
mantle transition zone due to the resistance resulting from positive buoyancy rendered by mineral phase
changes and a viscosity jump at the 660-km discontinuity (Chen et al., 2017; Fukao & Obayashi, 2013;
Huang & Zhao, 2006; Wei et al., 2012; Zhao, 2004). Seismic tomography and electrical conductivity data favor
a hydrated upper mantle existing above the flat slab, forming a big mantle wedge (BMW; Ichiki et al., 2006;
Kameyama & Nishioka, 2012; Zhao, 2004) because of the subduction driven corner flow and fluids from deep
slab dehydration and/or fluids brought down from the shallow mantle wedge by convection (Zhao et al.,
2011). The hot and wet upwelling in the BMW is considered to be responsible for the extensive Late
Mesozoic-Cenozoic tectono-magmatism in East Asia, causing significant intraplate volcanoes and earth-
©2018. American Geophysical Union. quakes, destruction of continental lithosphere, and a boundary in the surface topography (Kimura et al.,
All Rights Reserved. 2018; Liu et al., 2017; Zhao et al., 2012).

LI ET AL. 3286
Geochemistry, Geophysics, Geosystems 10.1029/2018GC007515

Figure 1. Topography, tectonic elements, and fault systems in the Shandong Province. Tectonic boundaries (modified from
Song, 2008) are shown in green dashed lines, and major faults are depicted with blue lines. The red regions show the
distribution of Cenozoic basalts in Shandong, after Shandong Provincial Bureau of Geology and Mineral Resource (1991).
The black stars denote two large ancient earthquakes (M ≈ 7.0, Anqiu earthquake, B.C. 70; M ≈ 8.5, Tancheng earthquake,
1668) in our study region. The inset shows the tectonic setting of the study region in eastern China depicted as the red
box. Abbreviations are as follows: CAOB, Central Asian Orogenic Belt; NCC, North China Craton; YZC, Yangtze Craton; TLFZ,
Tanlu Fault Zone; PQFZ, Penglai-Qixia Fault Zone; MJFZ, Muping-Jimo Fault Zone.

Although these studies present conceptual models for eastern China intraplate volcanism, they do not
address specifically why the volcanism is focused under isolated regions and not distributed along strike
(Faccenna et al., 2010; Tang et al., 2014). High-resolution seismic images reveal heterogeneities inside the
BMW. Seismic tomographic images of northeastern China obtained from a densely distributed seismic net-
work indicate that the hot mantle materials are more centralized beneath the Changbai volcanoes (Chen
et al., 2017; Zhao & Tian, 2013). In Shandong province, Mesozoic volcanism is widely spread. However, the
Cenozoic volcanism is only sparsely found in the east Changle-Linqu and Penglai regions (Zeng et al.,
2010, 2011; Xu et al., 2012), which are located at the northern segment of the Yishu fault zone (YSFZ) and
Jiaobei uplift region (Figure 1), respectively. Whether localized focused mantle upwelling exists beneath
these intraplate Cenozoic basalts remains an unresolved question.
The Tanlu Fault Zone, the main active strike-slip fault zone in eastern China, is believed to have facilitated the
upwelling of asthenosphere and played an important role in the Mesozoic-Cenozoic thinning of the cratonic
lithosphere (Liu & Niu, 2011; Xu et al., 2004). The central segment of the Tanlu Fault Zone in the Shandong
area, hereinafter referred to as the YSFZ, is an important mobile tectonic zone in eastern China. Resolving
the crustal/upper-mantle structural variations for the YSFZ and its surrounding areas has important implica-
tions for understanding the structural evolution and geodynamics in eastern China.
In this study, we use long-term ambient seismic noise data to construct Rayleigh wave phase-velocity
maps in the period band 5–30 s and invert these for high-resolution 3-D shear wave velocity (Vs) variations
in the crust and upper mantle beneath the YSFZ and surrounding areas. Our velocity model reveals fine-
scale crustal and uppermost mantle structure beneath the regions of Cenozoic volcanism, and the results

LI ET AL. 3287
Geochemistry, Geophysics, Geosystems 10.1029/2018GC007515

shed new light on the origin and dynamic processes of the intraplate
volcanism in the Shandong region, as well as the eastern China
continental margin.

2. Data and Analysis


Continuous seismic ambient noise data recorded at 47 broadband seismic
stations (Figure 2) are used in this study for imaging structure of the litho-
sphere beneath the Shandong region. The data were collected from
January 2014 to January 2015 with a sampling rate of 100 Hz. We estimate
the interstation Rayleigh wave phase-velocity dispersion for all available
station pairs from empirical Green’s functions (EGFs), which are con-
structed from long-time noise cross-correlation functions (NCFs) following
Yao et al. (2006, 2010). Only the vertical component recordings are consid-
ered to extract Rayleigh wave phase velocities for this study. The data pro-
cessing procedure is as follows: (1) The continuous raw time series data
Figure 2. Station distribution and raypath coverage. were divided into daily records and downsampled to 10 Hz for each sta-
tion. (2) The mean, linear trend, and instrument responses were removed
from each recording, and time domain normalization was applied using the running-absolute-mean method
(Bensen et al., 2007) to eliminate unwanted earthquake signals and instrument irregularities that can contam-
inate the seismic ambient noise signal. (3) Spectral whitening was applied to broaden the frequency band of
the ambient noise data, and a band-pass filter between 0.02 and 1 Hz was used to remove unwanted high-
and low-frequency signals. (4) The daily recordings from all station pairs were cross correlated, and the result-
ing cross-correlation functions were stacked to obtain the NCFs. The resulting NCFs calculated for all station
pairs show fairly good symmetry (Figure 3), indicating a similar energy flux from all directions into the array.
The EGFs are estimated by taking the Hilbert transform of the NCFs.
After the EGFs are computed for the Z-Z components of every station pair, several selection criteria are
applied prior to tomographic inversion. First, the interstation distance must be longer than 2 wavelengths
to satisfy the far-field approximation. Most previous time domain dispersion measurements from CFs or
EGFs apply a minimum interstation distance of 3 wavelengths (Bensen et al., 2007; Lin et al., 2009;
Porritt et al., 2011; Yao et al., 2006). However, Yao et al. (2011) applied a minimum interstation distance of
1.5 wavelengths to obtain dispersion at longer periods for relatively short interstation distances, which
may help resolve deeper structure. Second, the signal-to-noise ratio (SNR) must be higher than 5 at the period
of interest, where SNR is defined as the maximum amplitude of the envelope around periods in the signal
window divided by the mean amplitude of the envelope of a 150-s long noise window immediately after
signal window. This quality control metric is less strict than applied in some other studies (Bensen et al.,
2007; Lin et al., 2009; Porritt et al., 2011). The lower SNR threshold applied in this study helps increase the lim-
ited number of available paths. An illustration of extracting phase-velocity dispersion curves using the image
transformation technique (Yao et al., 2006) appears in Figure 4. The Rayleigh wave dispersion curves within
the period band 1–40 s for all useable station pairs are shown in Figure 5a, and the number of paths at each
period is shown in Figure 5b. The path coverage of Rayleigh wave phase-velocity measurements at six repre-
sentative periods, shown in Figure S1 in the supporting information, is good at periods from 10 to 25 s but
somewhat sparser at 5 and 30 s.

3. Phase Velocity Maps


We use the continuous regionalization following Montagner (1986) and Yao et al. (2010) and the generalized
inversion scheme of Tarantola and Valette (1982) to invert path-averaged phase velocities at each period and
produce 2-D phase-velocity variations (i.e., phase-velocity maps). The inversion for the isotropic phase velo-
city is controlled by three parameters: the standard error of the phase-velocity measurements, the a priori
parameter error, and the spatial correlation length. Previous studies of phase-velocity measurements (Yao
et al., 2010) suggest that the standard error of phase-velocity measurements is about 1%–2%, and we set
it to 2% for all measurements for this study. The choices for the a priori parameter error and the spatial
correlation length are somewhat subjective and are determined following Yao et al. (2010). The a priori

LI ET AL. 3288
Geochemistry, Geophysics, Geosystems 10.1029/2018GC007515

Figure 3. Cross-correlation functions in the period band 5–30 s as a function of interstation distance and lag time for all
station pairs.

parameter error, which constrains the anomaly amplitude, is set to be twice the standard deviation of all
observed phase velocities at each period. The spatial correlation length, which constrains the smoothness
of the model parameters, depends on raypath coverage at each period. The minimum value is usually
larger than one-third wavelength. We set it to be about 35–50 km.
Before analyzing the inversion phase velocity results, we first conduct checkerboard resolution tests to eval-
uate the adequacy of ray coverage and reliability of the inversion results according to Griot et al. (1998) and
Simons et al. (2002). In checkerboard tests, alternative positive and negative velocity anomalies (±8%) were
assigned to the 3-D grid nodes in the modeling space (Figure 6). Then we calculate synthetic dispersion
curves between stations at various periods for the same paths used in the real tomography. The synthetic dis-
persion curves are inverted to estimate the phase-velocity at each grid point. The recovered phase-velocity
maps for a 1° × 1° checkerboard model at periods 5, 10, 15, 20, 25, and 30 s are shown in Figure 6. The check-
erboard pattern and the amplitude of the input velocity perturbation are almost fully recovered beneath the
seismic array area, especially for the regions along the YSFZ. The edge of the study area is not well recovered,
due to lack of station pairs (Figure S1). Different checkerboard tests with different grid spacings are helpful to
evaluate the spatial resolution of the recovered model. Another checkerboard test with 0.75° × 0.75° grid spa-
cing is also conducted (Figure S2), revealing that the input model could be recovered in most parts of the
study area including the northern YSFZ and PQFZ. The horizontal resolution of isotropic phase velocity varia-
tions is about 80 km in the array area (Figure S2).
The phase-velocity maps estimated from the measured NCFs at 5, 10, 15, 20, 25, and 30 s are shown in
Figures 7a–7f. The image at 5 s correlates well with the major surface geological features; that is, low-velocity
anomalies appear in the sedimentary areas of the Jiaolai basin, whereas high-velocity anomalies appear in
mountainous regions. A N-S trending low-velocity anomaly in the crust beneath the northern YSFZ is

LI ET AL. 3289
Geochemistry, Geophysics, Geosystems 10.1029/2018GC007515

Figure 4. Representative examples of extracted phase-velocity dispersion curves for station pairs using the image transfor-
mation technique. Red line: interstation distance > twice the wavelength. The light blue circles show high signal-to-noise
ratio (>5) dispersion data, and the red dots indicate data used in the analysis.

clearly imaged on the phase-velocity map at 10 s (Figure 7b). In general, the phase-velocity structures of the
YSFZ present clear tectonic segmentation characteristics, especially for the eastern flank of the northern YSFZ
that correlates well with the velocity boundaries. Prominent low-velocity anomalies exist in the northern part
of the YSFZ at 5, 10, 15, 25, and 30 s, distinct from the southern part of the YSFZ, where relatively high-velocity
anomalies are observed (Figures 7a–c and 7e and 7f). The velocity anomaly boundary at ~36.5°N along the
YSFZ correlates well with the epicenter of the large historical Anqiu earthquake. The faults at the Jiaobei
uplift are also constrained by the velocity contrasts on either sides of them (Figures 7c–7f). In particular,
Figures 7e and 7f show low-velocity anomalies beneath the northern YSFZ and the Penglai areas, which
are spatially coincident with locations of the exposed Cenozoic basalts.

Figure 5. (a) All interstation dispersion curves from the time-frequency analysis (black lines) and the average dispersion
curve (red line). (b) Number of paths for different periods.

LI ET AL. 3290
Geochemistry, Geophysics, Geosystems 10.1029/2018GC007515

Figure 6. Checkerboard tests for resolution. 1° × 1° isotropic phase-velocity checkerboard model is shown in the upper left
panel. (a–f) Recovery isotropic phase-velocity maps at six periods (5, 10, 15, 20, 25, and 30 s).

4. Crustal and Upper Mantle Structure


Phase-velocity dispersion curves at each spatial location for 5- to 30-s period are extracted from the phase-
velocity maps and are inverted for the Vs profile using the surf96 program in CPS (Herrmann, 2013), with
results illustrated in Figure 8. The inversion is parameterized as a 28-layer smooth model with 2-km-thick

LI ET AL. 3291
Geochemistry, Geophysics, Geosystems 10.1029/2018GC007515

Figure 7. (a–f) Phase-velocity distributions for periods indicated on panels. The color bar at the left of each panel shows the
phase-velocity (km/s) range.

layers in the upper 10 km, 5-km-thick layers between 20 and 50-km depth, and 10-km-thick layers below 50-
km depth. This parameterization is chosen to provide relatively fine-scale layering where depth resolution is
relatively high and broader-scale layering where resolution is lower (Porritt et al., 2016). The starting model is
set as continental AK135 in the mantle, but with a constant velocity of ~4.5 km/s in the upper ~130 km of the
model to avoid artificially inserting low-velocity layers within the crust (Herrmann, 2013; black line in Figure 8b).
The inversion is damped toward small, smooth changes in velocity, and a total of 30 iterations are run at
each grid point. The observed dispersion curve and its predication from the estimated model (red line in
Figure 8b) at a grid point (118.5°E, 35.5°N) are shown in Figure 8a. We observe a good fit between the
observed and predicted dispersion curves. The sensitivity kernels as a function of depth are shown in
Figure 8c at period of 5, 10, 15, 20, 25, and 30 s, which shows the decrease of resolution with depth.
Rayleigh wave phase-velocity dispersion at 25 and 30 s are still a little sensitive to depths of 50–80 km and
represent the features of upper mantle.

4.1. Shear-Wave Velocity Structure


The lateral variation of Vs at depths of 5, 10, 20, 25, 30, 35, 50, 60, and 80 km is depicted in Figure 9. Figure 10
shows the Vs heterogeneity along six vertical profile lines from the surface to 80-km depth. Representative
examples of shear wave velocity versus depth for different geographic and tectonic area are shown in
Figure S3.
At shallow depth (5 km, Figure 9a), lateral velocity variations correlate well with known near-surface geology.
Low Vs is observed in the Jiaolai basin, probably due to the presence of thick sedimentary cover. High velo-
cities are observed in the Jiaobei uplift except the PQFZ. The Luxi uplift generally shows a small horizontal
velocity variation. The shear wave structure is clearly dominated by the contrast between the high velocity
in the northern YSFZ and lower velocity in the southern YSFZ.

LI ET AL. 3292
Geochemistry, Geophysics, Geosystems 10.1029/2018GC007515

Figure 8. (a) Observed and predicted dispersion curves. (b) 1-D Vs model from inversion and starting model. (c) Phase-velocity sensitivity kernels as a function of
depth at the periods indicated.

In the upper/middle crust (Figures 9b and 9c), the Vs pattern is similar between 10- and 20-km depth, which is
generally reversed compared to that at 5-km depth (Figure 9a). Note that the northern YSFZ appears as low
velocities, which are also observed in the vertical cross section through the northern YSFZ (Figures 10a–10d).
The Vs pattern is consistent with the phase velocity map at 10 s (Figure 7b).
In the lower crust (25-, 30-, and 35-km depth in Figures 9d–9f, respectively), the lateral velocity variation in the
study area becomes larger with increasing depth. More specifically, the maximum difference between high
velocity and low velocity is about 0.4 km/s at 25 km (Figure 9d), 0.6 km/s at 30 km (Figure 9e), and 0.8 km/s
at 35 km (Figure 9f). These variations are also revealed from the cross sections (Figure 10). High velocities
are observed beneath the northern YSFZ and relatively low velocities beneath the southern YSFZ. The
PQFZ in the Jiaobei uplift is marked by relatively high velocity. The Luxi uplift shows a relatively low velocity
at 35-km depth, with Vs generally about 3.9 km/s.
In the upper mantle (50-, 60-, and 80-km depth in Figures 9g–9i, respectively), the Vs pattern is similar except
for a relatively high-velocity anomaly beneath the northern of YSFZ. The low-velocity zone (LVZ) appears
obviously beneath the northern YSFZ and the PQFZ of the Jiaobei uplift, respectively, which is also observed
from the vertical cross-section maps (Figure 10). The scale and amplitude of these low-velocity anomalies
increase with depth, extending to at least 80-km depth. These low-velocity anomalies are coincident with
the Moho uplift and LAB apex imaged by receiver function studies (Chen et al., 2006). In addition, our imaged
low-velocity features have a 3–10% velocity drop compared to the surrounding mantle, which is also highly
consistent with previous studies (Chen et al., 2006).
Our 3-D model reveals two LVZs below 50-km depth, in the upper mantle beneath the northern YSFZ and the
Penglai area, where isolated exposed Cenozoic basalts are found. The reliability of these LVZs is critical to the
following discussion on the interaction between the upper mantle material upwelling, crustal thinning pro-
cesses, and intraplate volcanism. In order to evaluate the reliability of these two LVZs features, we conduct
detailed synthetic resolution tests and additional analysis.

LI ET AL. 3293
Geochemistry, Geophysics, Geosystems 10.1029/2018GC007515

Figure 9. (a–i) Lateral variation of Vs at the depths indicated. The black stars in (b) and (c) denote the epicenters of the historical Anqiu (M7.0) earthquake. The color
bars indicate the Vs scale for each panel (km/s).

The most sensitive depth for the Rayleigh wave with a period of 30 s is around 40 km (Figure 8c) in view of the
fact that the physical resolution is limited by the predominant wavelengths considered. The kernel sensitivity
values for periods of 25 and 30 s show that it still has resolution at depths of 50–80 km. This indicates that the
Rayleigh wave phase-velocity dispersions used in this study can constrain Vs structure at 50- to 80-km depth.
We conduct synthetic tests to evaluate the reliability of the major estimated velocity features. The synthetic
tests are carried out following four steps: (1) We first construct a 3-D input model with two uppermost mantle
LVZ features similar to the estimated 3-D model. The Vs perturbation values of two LVZs are set to 6% at 50-
km depth and 8% at 60- to 80-km depth (Figure S4). (2) We calculate synthetic dispersion curves between
stations at 5- to 30-s periods for the same paths used in the real tomography. (3) The synthetic phase-velocity
maps are estimated from the synthetic dispersion data with 2% additive Gaussian white noise by tomo-
graphic inversion. (4) The local dispersion curves are extracted and inverted to estimate the shear wave velo-
city profile at each grid point. Figure S4 shows the input and recovered model at different depths, indicating
that the low-velocity features at 50- to 80-km depths could be reconstructed well.
We also conduct one additional synthetic test to confirm the reliability of other velocity features. The four
steps are the same as above, and the only difference is that the input model (Figure 9) in this test is the whole
estimated model including all the velocity anomalies, instead of only the two uppermost mantle LVZ features.
Figure S5 shows the input model and the recovered model at the six grid points, which are selected from the
area of the LVZs. Although the velocity structures of the input model can be essentially reconstructed, the
resolution is relatively lower at depths below 50 km. The maximum error value is about 4.8% at depths below
50 km. The minimum Vs perturbation value of the low-velocity anomalies revealed in our final velocity model
below 50 km is about 5.6%, which is still larger than the maximum inversion error (4.8%). This implies that our
data set is able to recover the significant low-velocity anomalies within the uppermost mantle. Figures S6
and S7 show the recovered model at different depths and along six vertical profiles, indicating that the Vs

LI ET AL. 3294
Geochemistry, Geophysics, Geosystems 10.1029/2018GC007515

features including the low-velocity features at 50- to 80-km depths are all
fairly well resolved. Our resolution-test results demonstrate that the two
LVZs below 50 km are reliable features, supporting the main conclusion
of this work.
In order to evaluate the robustness of the uppermost mantle low-
velocity anomalies, we also compared our result to previously regional
P wave tomography studies. The regional P wave tomographic images
of Wang et al. (2014; Figure S8) show that a prominent low-velocity
anomaly appears beneath the northern YSFZ and the adjacent Bohai
Basin and Shandong Penisula extending to a depth of 200 km. The upper
mantle LVZs in our model correlates well with this low-velocity anomaly.
Hence, we think that the velocity structure below 50 km reveals a reliable
pattern of velocity contrasts between the two LVZs and the surrounding
regions, though the amplitude of the velocity perturbation may not be
estimated precisely.

4.2. Variation of Moho Depth


Previous deep seismic sounding profile studies (Jiang et al., 2000; Jia &
Zhang, 2005; Zhang et al., 1996) show that the depth of Moho interface
is 31–37 km and the P wave velocity of the lower crust is 6.5–6.9 km/s
in the Shandong area. Following Su et al. (2016), we choose the
average P wave velocity of the lower crust in the study area as the
P wave velocity of the Moho interface (6.7 km/s). The corresponding
shear wave velocity is about 3.87 km/s, considering the relationship
between P and S wave velocities: Vp/Vs = 1.732. In this work, the
Moho interface is defined as the crust-mantle boundary at Vs = 3.87 km/s.
The value of Moho depth is used to illustrate the approximate
crustal thickness.
The lateral heterogeneities in Moho depth beneath the seismic array are
displayed in Figure 11. The imaged Moho depth exhibits significant lateral
undulation, from 32–37 km in western Shandong to 25–34 km in eastern
Shandong. From south to north along the YSFZ, the Moho depth
Figure 10. (a–f) Vertical cross sections of isotropic Vs, with the profile loca-
decreases gradually from 34 to 25 km. The Moho beneath the northern
tions shown in the bottom panel. Topography is depicted above each pro-
file as the black area. The red triangles denote the locations of Cenozoic segment of YSFZ is shallower than in the surrounding regions
basalts, and the two black stars denote the epicenters of the historical Anqiu (Figure 11). The crust beneath the northern YSFZ is thinned by ~5 km
(M7.0) and Tancheng (M8.5) earthquakes. Note that the focal depth of the (Figures 10a–10d and 11). Our results are highly consistent with the
historical earthquake is unknown. The dashed line in (a) is the Moho depth Moho depths obtained from receiver function studies (Chen et al., 2006;
constructed from receiver-function migration studies (Chen et al., 2006).
Figure 10a). The Moho depth and therefore crustal thickness vary from
~28 to 33 km beneath the Jiaobei uplift. The Moho depth beneath the
PQFZ is obviously shallower (Figures 10b and 10f, and 11).

5. Discussion
5.1. Implications of Crustal Velocity Anomalies
The crustal velocity structures within the Luxi uplift and Sulu orogen are very homogeneous, without
significant velocity perturbations. The southern YSFZ is not acting as a sharp velocity boundary
between the Luxi uplift and Sulu orogen. Instead, it shows similar character to regions on both sides
(Figures 7a–7f). The reason for the homogenous crustal velocity might be that the Luxi Terrane and Sulu
Orogen are old and stable, barely affected by Cenozoic tectonic processes (Deng et al., 2018). By contrast,
the northern YSFZ correlates well with the boundary of high-velocity and low-velocity anomalies extending
from the surface to the upper mantle. This implies that the northern YSFZ may have been reactivated
during the Cenozoic period, when continental extension and crustal thinning processes occurred along
the east China margin.

LI ET AL. 3295
Geochemistry, Geophysics, Geosystems 10.1029/2018GC007515

Figure 11. Variations of Moho depth (corresponding to Vs = 3.87 km/s) over the study area. The red triangles denote the
locations of Cenozoic basalts, and the two black stars denote the epicenters of historical Anqiu (M7.0) and Tancheng (M8.5)
earthquakes.

An LVZ appears in the upper-middle crust beneath the northern YSFZ (Figures 9b and 9c and 10a–10d), while
relatively high-velocity anomalies appear on each side, which is consistent with previous results (Li et al.,
2006; Su et al., 2016; Tian et al., 2009; Zhang et al., 1996). This LVZ is spatially consistent with a zone of high
conductivity (Hu et al., 2014), implying that the LVZ may be hydrated due to water filling along the preexisted
fault zone. Another possible explanation for this LVZ is that it may be related to magmatic dike intrusion from
underlying fluid-rich magma. In either explanation, this LVZ should act as a viscously weak zone embedding
in the middle crust below the northern YSFZ, being apt to concentrate tectonic stress (Chen et al., 2014;
Kenner & Segall, 2000). It may explain why the epicenter of the historical M7.0 Anqiu earthquake is generally
located close to the southern boundary of the crustal LVZ (Figures 9b and 9c). Fluid- and/or melt-related LVZs
are also found in source areas of many large crustal earthquakes (e.g., Chen et al., 2014; Cheng et al., 2011;
Zhao et al., 2002).
The upper-middle crust LVZ beneath the northern YSFZ does not extend to the lower crust. Instead, a high-
velocity anomaly appears (Figures 9e and 9f). There are two possible explanations for this high-velocity zone
(HVZ). One is that the HVZ corresponds to juvenile lower crust resulting from the basalt underplating at the
bottom of the Moho. Similar lower-crust layers are also found in other extensional geological settings, for
instance, the rift continental margin along the northern South China Sea (Wan et al., 2017) and the active
Baikal rift zone (Thybo & Nielsen, 2009). The other explanation is that the HVZ is composed of the upwelling
lithospheric mantle resulting from the continental extension and subsequent crustal thinning. The tectonic
background and geodynamic processes related to the two explanations are similar. The difference between
these explanations is that the first considers the nature of this HVZ as juvenile lower crust, while the second
considers the nature of this HVZ as uppermost mantle. The main determinant to distinguish between these
two explanations is whether the Moho is flat or uplifted. Previous studies such as the thermal structure of the
lithosphere across the YSFZ (Lai et al., 2007; Zu et al., 1996), the Moho discontinuities topography using recei-
ver function methods (Chen et al., 2006), and wide-angle seismic profiles (Jia et al., 2014; Li et al., 2011; Xia
et al., 2017) indicate that the YSFZ has the features of mantle uplift and crustal thinning. Hence, in this work,
we prefer the second explanation; that is, the YSFZ is a conventional rift zone including three characteristic
features: surface manifestation as a low-lying topographic trough, Moho uplift due to crustal thinning, and
reduced seismic velocity in the uppermost mantle due to decompression melting or heating from the
Earth’s interior (Thybo & Nielsen, 2009).

LI ET AL. 3296
Geochemistry, Geophysics, Geosystems 10.1029/2018GC007515

Figure 12. Three-dimensional view of the main features of the velocity structure. The red cones at the surface denote the
locations of Cenozoic basalts. The red iso-volume of low-velocity anomalies at the bottom is rendered where Vs pertur-
bations are less than 5%. Note the spatial correlation between the focused mantle upwelling in the uppermost mantle,
the regions of crustal thinning, and the Cenozoic basalts.

5.2. Focused Mantle Upwelling and Cenozoic Intraplate Volcanism


First-order evidence from previous low-resolution P wave tomography studies suggested the presence of
wide-spread Cenozoic basalts in the eastern China continental margin (Sakuyama et al., 2013; Zhao et al.,
2011). The studies associated the origin of intraplate volcanoes in the region with the hot and wet mantle
upwelling (Huang & Zhao, 2009; Wei et al., 2012). The densely distributed local and temporary seismic arrays
allow us to resolve the finer-scale structures in this region. Our results show that the basalts are characterized
by low Vs anomalies at ~60- to 80-km depth and isolated and distributed in focused regions (Figures 10 and
12.). In the Shandong region, Cenozoic basalts are mainly distributed under the Changle-Linqu area in
north-central Shandong, while some are found in the Jiaobei uplift (Zeng et al., 2011). The main areas of out-
cropping Cenozoic basalts are spatially close to the northern YSFZ and PQFZ.
The imaged Moho depth values are significantly lower beneath the northern YSFZ and the PQFZ of the
Jiaobei uplift, which is consistent with previous studies (Chen et al., 2006; Jia & Zhang, 2005; Su et al., 2016;
Zhang et al., 1996). Moho depths derived from gravity data show that uplift of the Moho occurs along the
whole YSFZ (Wang et al., 2015). However, our images only show Moho uplift at the northern segment of
the YSFZ, instead of the whole YSFZ. Given that the Cenozoic basalts only occur along the northern segment
of the YSFZ, our high-resolution Vs images show more detailed lithospheric structures than previous work.
The Moho tilts from east to west with an obvious uplift beneath the YSFZ, suggesting that this fault zone
is a major boundary within Shandong region. The coincidence of the Moho uplift with the surface location
of the northern YSFZ (Figures 10a–10e and 11) provides evidence for strike-slip extensional structure and
its deep penetration into the lithospheric mantle (Xiong et al., 2011). An arc-shaped mantle upwelling at
~60- to 80-km depth beneath the northern YSFZ and the PQFZ is revealed in the images (labeled LVZ in
Figures 10a and 10b), and its apex is roughly coincident with the location of uplift. Our images support the
arguments that the YSFZ has facilitated the upwelling of hot asthenosphere materials during late Cenozoic
continental extension and lithospheric thinning in eastern China.
The asthenospheric upwelling corresponding with the low-velocity anomalies could be attributable to a high
temperature gradient, the presence of melt in the asthenosphere, and compositional contrasts between the
cold, refractory lithospheric remnant, and the uplifted hot, fertile asthenosphere (Conrad et al., 2010;

LI ET AL. 3297
Geochemistry, Geophysics, Geosystems 10.1029/2018GC007515

Hammond et al., 2013). The focused mantle upwelling in our research area most likely corresponds to partial
melting caused by decompression melting beneath the thinned crust. The low velocities in the uppermost
mantle are linked to melt that feeds volcanism. It is clear that mantle material upwells beneath the northern
YSFZ and the Jiaobei uplift (Figures 10, 11, and 12.). The strong spatial correlation between the narrow low Vs
zones in the uppermost mantle and the thinned crust, as well as the locations of the isolated accumulated
volcanisms, suggests that centrally fed magmatic segments are supplied by localized asthenospheric
upwelling. Thermal modeling studies (Liu et al., 2016) show that the northern YSFZ exhibits significantly high
surface heat flow of more than 60 mWm 2 and a high proportion of mantle to surface heat flow of 60%. This
result indicates that the northern YSFZ has a “cold crust but hot mantle” structure, which may explain the low-
velocity features in the uppermost mantle beneath the northern YSFZ.
The broad upper-mantle regions above the stagnant Pacific slab have formed a BMW under NE China. The
intraplate back-arc magmatism and volcanism in eastern China are caused by the deep dehydration process
of the subducting slab and the convective process of the mantle wedge (Stern, 2002; Zhao, 2012; Zhao et al.,
2007). Recent studies of dyke swarms in the western Jiaodong Peninsula also suggest that these rocks were
derived from lithospheric and asthenospheric sources and that the magma tectonics are related to rapid
lithospheric thinning (Li, Li, et al., 2016; Ma, Jiang, Hofmann, et al., 2014; Ma, Jiang, Hou, et al., 2014). The
BMW may represent the large-scale upwelling of hot asthenospheric materials, leading to the formation of
the continental rift systems as well as intraplate volcanoes in Northeast Asia. After asthenospheric materials
originate from the stagnant slab in the mantle transition zone, they rise to shallower depth and are focused
into smaller features at 50- to 80-km depth. These focused features may suggest that the onset of centrally
fed mantle upwelling contributed to isolated focused second-order intraplate volcanism. The LVZ in our
images may correspond to zones of increased melt production or reduced density of depleted mantle.
According to GPS studies (Yin et al., 2008), the whole Shandong block is moving southeastward, which causes
the magma to flow westward. This may explain the phenomena in our model that the LVZ in the uppermost
mantle does not directly connect to the lower crust but migrates westward (Figures 10a and 10b).
In addition to north-central Shandong, Cenozoic basalts in the NCC are also found in east-central Liaoning,
eastern Jilin, northwest Hebei, and eastern and northern Shanxi (Chen et al., 2015; Li, Ma, & Robinson,
2016; Sakuyama et al., 2013). Lithospheric thinning has long been proposed as the dominant process occur-
ring in all of eastern China during the Cenozoic (Chen et al., 2006; Xu et al., 2012). We speculate that the spa-
tial correlation between the locations of Cenozoic basalts, the region of Moho uplift, and the low velocities, as
imaged in Shandong region in this study (Figure 12.), represents a typical tectonic model of Cenozoic intra-
plate volcanism. Localized low-velocity anomalies are also revealed clearly in the crust and the BMW directly
beneath the active Changbai and Ulleung volcanoes (Chen et al., 2017). Similar relationships between the
focused diapiric mantle upwelling, crustal rifting segmentation, and surface volcanisms are also found in
the Afar Depression, at the northern end of the East African Rift, where the transition from plume-induced
continental breakup to seafloor spreading is active today. Regularly spaced, focused, LVZs, interpreted as
upwellings, are also seen beneath the youthful Gulf of California (Mexico) plate boundary (Wang et al., 2009).
Besides the intraplate volcanism during the Cenozoic era, the eastern part of the North China Craton, includ-
ing the Jiaodong region, had also experienced large-scale lithosphere thinning, extensive magmatic intru-
sion, and volcanic eruptions during the Mesozoic Yanshanian orogeny (Bunge & Grand, 2000; Zhang et al.,
2010; Liu et al., 2017). Since our tomographic images cannot reveal the ancient status of the Earth during
the Early Cretaceous (~140–110 Ma), we only discuss the origins of the Cenozoic intraplate volcanisms in this
paper. There is no clear spatial connection between the location of the current focused mantle upwelling and
the widely spread ancient Mesozoic basalts in eastern China. This inconsistency may imply that the mantle
structure has changed since the Early Cretaceous, when the North China Craton was destroyed and intraplate
volcanism occurred resulting from the subduction of the Izanagi (or Paleo-Pacific) plate.

6. Conclusions
High-resolution Rayleigh wave phase velocity maps in the period band 5–30 s are constructed from the cross
correlation of ambient noise recorded by 47 broadband stations in the Shandong province of China. The crust
and upper mantle structures beneath the Shandong region of East China are well imaged from these phase
velocity maps. Our results suggest that there are significant differences between southern and northern

LI ET AL. 3298
Geochemistry, Geophysics, Geosystems 10.1029/2018GC007515

portions of the YSFZ, with prominent LVZs in the crust and asthenosphere beneath the northern portion.
Moho depths show small-scale undulations beneath west and east Shandong and a significant uplift from
south to north beneath the YSFZ.
The coincidence of the Moho uplift and the apex of the LVZ at the asthenosphere with the surface location
of the northern YSFZ indicates that the YSFZ is a large, deep fault zone on the lithospheric scale with strike-
slip extensional structure and that the fault zone serves as a major channel for asthenosphere upwelling,
playing an important role in the Cenozoic continental extension and lithospheric thinning and
intraplate volcanism.
The prominent LVZ observed in the asthenosphere beneath the northern of the YSFZ and the Jiaobei uplift
has a strong spatial coincidence with Moho uplift, as well as the locations of the isolated Cenozoic volcanisms
in our research area. This consistency suggests that the asthenosphere upwelling induced by the stagnant
Pacific slab is focused into localized smaller features and supplies the Cenozoic volcanoes of the eastern
China continental margin.

Acknowledgments References
We would like to thank the Seismic
Network Center at the Earthquake Bensen, G. D., Ritzwoller, M. H., Barmin, M. P., Levshin, A. L., Lin, F., Moschetti, M. P., et al. (2007). Processing seismic ambient noise data to
Administration of Shandong Province obtain reliable broad-band surface wave dispersion measurements. Geophysical Journal International, 169(3), 1239–1260. https://doi.org/
for providing the waveform data. The 10.1111/j.1365-246X.2007.03374.x
waveform data used for this study are Bunge, H.-P., & Grand, S. P. (2000). Mesozoic plate-motion history below the Northeast Pacific Ocean from seismic images of the subducted
archived on a webpage: https://pan.baidu. Farallon slab. Nature, 405(6784), 337–340. https://doi.org/10.1038/35012586
com/s/1V1iNYwO9xJ1lu55CSp93CQ. We Chen, C., Zhao, D., Tian, Y., Wu, S., Hasegawa, A., Lei, J., et al. (2017). Mantle transition zone, stagnant slab and intraplate volcanism in
also thank the editor, Maureen Long, Northeast Asia. Geophysical Journal International, 209(1), ggw491–ggw485. https://doi.org/10.1093/gji/ggw491
and two anonymous reviewers for Chen, C., Zhao, D., & Wu, S. (2014). Crust and upper mantle structure of the New Madrid seismic zone: Insight into intraplate earthquakes.
their constructive comments and Physics of the Earth and Planetary Interiors, 230, 1–14. https://doi.org/10.1016/j.pepi.2014.01.016
suggestions, which significantly Chen, H., Xia, Q.-K., Ingrin, J., Jia, Z.-B., & Feng, M. (2015). Changing recycled oceanic components in the mantle source of the Shuangliao
improved the quality of this paper. This Cenozoic basalts, NE China: New constraints from water content. Tectonophysics, 650, 113–123. https://doi.org/10.1016/j.
work was supported by the Laboratory tecto.2014.07.022
for Marine Geology, Qingdao National Chen, L., Zheng, T., & Xu, W. (2006). A thinned lithospheric image of the Tanlu Fault Zone, eastern China: Constructed from wave equation
Laboratory for Marine Science and based receiver function migration. Journal of Geophysical Research, 111, B09312. https://doi.org/10.1029/2005JB003974
Technology (grant MGQNLM- Cheng, B., Zhao, D., & Zhang, G. (2011). Seismic tomography and anisotropy in the source area of the 2008 Iwate-Miyagi earthquake (M 7.2).
KF201706), the National Program on Physics of the Earth and Planetary Interiors, 184(3-4), 172–185. https://doi.org/10.1016/j.pepi.2010.11.006
Global Change and Air-Sea Interaction Conrad, C. P., Wu, B., Smith, E. I., Bianco, T. A., & Tibbetts, A. (2010). Shear-driven upwelling induced by lateral viscosity variations and asth-
(GASI-GEOGE-02), the Research enospheric shear: A mechanism for intraplate volcanism. Physics of the Earth and Planetary Interiors, 178(3-4), 162–175. https://doi.org/
Program of Earthquake Administration 10.1016/j.pepi.2009.10.001
of Shandong Province (JJ1601), and the Deng, J., Wang, C., Bagas, L., Santosh, M., & Yao, E. (2018). Crustal architecture and metallogenesis in the south-eastern North China Craton.
National Natural Science Foundation of Earth-Science Reviews, 182, 251–272. https://doi.org/10.1016/j.earscirev.2018.05.001
China (41606055). Faccenna, C., Becker, T. W., Lallemand, S., Lagabrielle, Y., Funiciello, F., & Piromallo, C. (2010). Subduction-triggered magmatic pulses: A new
class of plumes? Earth and Planetary Science Letters, 299(1–2), 54–68. https://doi.org/10.1016/j.epsl.2010.08.012
Fukao, Y., & Obayashi, M. (2013). Subducted slabs stagnant above, penetrating through, and trapped below the 660 km discontinuity. Journal
of Geophysical Research: Solid Earth, 118, 5920–5938. https://doi.org/10.1002/2013JB010466
Griot, D. A., Montagner, J. P., & Tapponnier, P. (1998). Phase velocity structure from Rayleigh and Love waves in Tibet and its neighboring
regions. Journal of Geophysical Research, 103(B9), 21,215–21,232. https://doi.org/10.1029/98JB00953
Hammond, J. O. S., Bastow, I. D., Kendall, J. M., Stuart, G. W., Wright, T. J., Ebinger, C. J., et al. (2013). Mantle upwelling and initiation of rift
segmentation beneath the Afar Depression. Geology, 41(6), 635–638. https://doi.org/10.1130/G33925.1
Herrmann, R. B. (2013). Computer programs in seismology: An evolving tool for instruction and research. Seismological Research Letters, 84(6),
1081–1088. https://doi.org/10.1785/0220110096
Hu, W., Zhu, G., Yan, L., & Zhan, R. (2014). Analysis of relationship between seismic activity and crust electrical textures for the central seg-
ment of the Tan-Lu Fault Zone. Geological Review, 60(1), 80–90.
Huang, J., & Zhao, D. (2006). High-resolution mantle tomography of China and surrounding regions. Journal of Geophysical Research, 111,
B09305. https://doi.org/10.1029/2005JB004066
Huang, J., & Zhao, D. (2009). Seismic imaging of the crust and upper mantle under Beijing and surrounding regions. Physics of the Earth and
Planetary Interiors, 173(3-4), 330–348. https://doi.org/10.1016/j.pepi.2009.01.015
Ichiki, M., Baba, K., Obayashi, M., & Utada, H. (2006). Water content and geotherm in the upper mantle above the stagnant slab: Interpretation
of electrical conductivity and seismic P-wave velocity models. Physics of the Earth and Planetary Interiors, 155(1-2), 1–15. https://doi.org/
10.1016/j.pepi.2005.09.010
Jia, S., Wang, F., Tian, X., Duan, Y., Zhang, J., Liu, B., & Lin, J. (2014). Crustal structure and tectonic study of North China Craton from a long deep
seismic sounding profile. Tectonophysics, 627, 48–56. https://doi.org/10.1016/j.tecto.2014.04.013
Jia, S., & Zhang, X. (2005). Crustal structure and comparison of different tectonic blocks in North China. Chinese Journal of Geophysics, 48(3),
611–620.
Jiang, W., Hao, T., Jiao, C., & Song, H. (2000). The characters of gravity and magnetic fields and crustal structure from Qingzhou to Muping,
Shandong Province. Progress in Geophysics, 15(4), 18–26.
Kameyama, M., & Nishioka, R. (2012). Generation of ascending flows in the Big Mantle Wedge (BMW) beneath northeast Asia induced by
retreat and stagnation of subducted slab. Geophysical Research Letters, 39(10), L10309. https://doi.org/10.1029/2012GL051678
Kenner, S. J., & Segall, P. (2000). A mechanical model for intraplate earthquakes: Application to the New Madrid seismic zone. Science,
289(5488), 2329–2332. https://doi.org/10.1126/science.289.5488.2329

LI ET AL. 3299
Geochemistry, Geophysics, Geosystems 10.1029/2018GC007515

Kimura, J.-I., Sakuyama, T., Miyazaki, T., Vaglarov, B. S., Fukao, Y., & Stern, R. J. (2018). Plume-stagnant slab-lithosphere interactions: Origin of
the late Cenozoic intra-plate basalts on the East Eurasia margin. Lithos, 300-301, 227–249. https://doi.org/10.1016/j.lithos.2017.12.003
Lai, X., Li, S., & Sun, Y. (2007). Deep tectonic background of three Ms>7.0 strong earthquakes in Bohai and its adjacent region. Journal of
Geodesy and Geodynamics, 27(1), 31–34.
Li, L., Li, S.-R., Santosh, M., Li, Q., Gu, Y., Lü, W.-J., et al. (2016). Dyke swarms and their role in the genesis of world-class gold deposits: Insights
from the Jiaodong peninsula, China. Journal of Asian Earth Sciences, 130, 2–22. https://doi.org/10.1016/j.jseaes.2016.06.015
Li, S., Lai, X., Liu, B., Wang, Z., He, J., & Sun, Y. (2011). Differences in lithospheric structures between two sides of Taihang Mountain obtained
from the Zhucheng-Yichuan deep seismic sounding profile. Science China Earth Sciences, 54(6), 871–880. https://doi.org/10.1007/s11430-
011-4191-4
Li, Y.-Q., Ma, C.-Q., & Robinson, P. T. (2016). Petrology and geochemistry of Cenozoic intra-plate basalts in east-central China: Constraints on
recycling of an oceanic slab in the source region. Lithos, 262(C), 27–43. https://doi.org/10.1016/j.lithos.2016.06.012
Li, Z., Xu, Y., Hao, T., Liu, J., & Zhang, L. (2006). Seismic tomography and velocity structure in the crust and upper mantle around Bohai Sea
area. Chinese Journal of Geophysics, 49(3), 797–804.
Lin, F.-C., Ritzwoller, M. H., & Snieder, R. (2009). Eikonal tomography: Surface wave tomography by phase front tracking across a regional
broad-band seismic array. Geophysical Journal International, 177(3), 1091–1110. https://doi.org/10.1111/j.1365-246X.2009.04105.x
Liu, H., & Niu, F. (2011). Receiver function study of the crustal structure of Northeast China: Seismic evidence for a mantle upwelling beneath
the eastern flank of the Songliao Basin and the Changbaishan region. Earthquake Science, 24(1), 27–33. https://doi.org/10.1007/s11589-
011-0766-6
Liu, Q., Zhang, L., Zhang, C., & He, L. (2016). Lithospheric thermal structure of the North China Craton and its geodynamic implications.
Journal of Geodynamics, 102, 139–150. https://doi.org/10.1016/j.jog.2016.09.005
Liu, X., Zhao, D., Li, S., & Wei, W. (2017). Age of the subducting Pacific slab beneath East Asia and its geodynamic implications. Earth and
Planetary Science Letters, 464, 166–174. https://doi.org/10.1016/j.epsl.2017.02.024
Ma, L., Jiang, S.-Y., Hofmann, A. W., Dai, B.-Z., Hou, M.-L., Zhao, K.-D., et al. (2014). Lithospheric and asthenospheric sources of lamprophyres in
the Jiaodong Peninsula: A consequence of rapid lithospheric thinning beneath the North China Craton? Geochimica et Cosmochimica
Acta, 124, 250–271. https://doi.org/10.1016/j.gca.2013.09.035
Ma, L., Jiang, S.-Y., Hou, M.-L., Dai, B.-Z., Jiang, Y.-H., Yang, T., et al. (2014). Geochemistry of Early Cretaceous calc-alkaline lamprophyres in the
Jiaodong Peninsula: Implication for lithospheric evolution of the eastern North China Craton. Gondwana Research, 25(2), 859–872. https://
doi.org/10.1016/j.gr.2013.05.012
Maruyama, S., Hasegawa, A., Santosh, M., Kogiso, T., Omori, S., Nakamura, H., et al. (2009). The dynamics of big mantle wedge, magma factory,
and metamorphic–metasomatic factory in subduction zones. Gondwana Research, 16(3–4), 414–430. https://doi.org/10.1016/j.
gr.2009.07.002
Montagner, J. -P. (1986). Regional three-dimensional structures using long period surface waves. Annales Geophysicae, 4, 283–294.
Porritt, R. W., Allen, R. M., Boyarko, D. C., & Brudzinski, M. R. (2011). Investigation of Cascadia segmentation with ambient noise tomography.
Earth and Planetary Science Letters, 309(1–2), 67–76. https://doi.org/10.1016/j.epsl.2011.06.026
Porritt, R. W., Miller, M. S., O’Driscoll, L. J., Harris, C. W., Roosmawati, N., & Teofilo da Costa, L. (2016). Continent-arc collision in the Banda
arc imaged by ambient noise tomography. Earth and Planetary Science Letters, 449, 246–258. https://doi.org/10.1016/j.
epsl.2016.06.011
Sakuyama, T., Tian, W., Kimura, J.-I., Fukao, Y., Hirahara, Y., Takahashi, T., et al. (2013). Melting of dehydrated oceanic crust from the stagnant
slab and of the hydrated mantle transition zone: Constraints from Cenozoic alkaline basalts in eastern China. Chemical Geology, 359, 32–48.
https://doi.org/10.1016/j.chemgeo.2013.09.012
Simons, F. J., van der Hilst, R. D., Montagner, J. P., & Zielhuis, A. (2002). Multimode Rayleigh wave inversion for heterogeneity and azimuthal
anisotropy of the Australian upper mantle. Geophysical Journal International, 151(3), 738–754. https://doi.org/10.1046/j.1365-
246X.2002.01787.x
Song, M. (2008). The composing, setting and evolution of tectonic units in Shandong Province. Geological Survey and Research, 31(3),
165–175.
Stern, R. J. (2002). Subduction zones. Reviews of Geophysics, 40(4), 1012. https://doi.org/10.1029/2001RG000108
Su, D., Fan, J., Wu, S., Chen, C., Dong, X., & Chen, S. (2016). 3D P wave velocity structures of crust and their relationship with earthquakes in the
Shandong area. Chinese Journal of Geophysics, 59(4), 1335–1349.
Tang, Y., Obayashi, M., Niu, F., Grand, S. P., Chen, Y. J., Kawakatsu, H., et al. (2014). Changbaishan volcanism in Northeast China linked to
subduction-induced mantle upwelling. Nature Geoscience, 7(6), 470–475. https://doi.org/10.1038/ngeo2166
Tang, Y. J., Zhang, H. F., Santosh, M., & Ying, J. F. (2013). Differential destruction of the North China Craton: A tectonic perspective. Journal of
Asian Earth Sciences, 78, 71–82. https://doi.org/10.1016/j.jseaes.2012.11.047
Tarantola, A., & Valette, B. (1982). Generalized nonlinear inverse problem solved using the least squares criterion. Reviews of Geophysics, 20(2),
219–232. https://doi.org/10.1029/RG020i002p00219
Thybo, H., & Nielsen, C. A. (2009). Magma-compensated crustal thinning in continental rift zones. Nature, 457(7231), 873–876. https://doi.org/
10.1038/nature07688
Tian, Y., Zhao, D., Sun, R., & Teng, J. (2009). Seismic imaging of the crust and upper mantle beneath the North China Craton. Physics of the
Earth and Planetary Interiors, 172(3–4), 169–182. https://doi.org/10.1016/j.pepi.2008.09.002
Wan, K., Xia, S., Cao, J., Sun, J., & Xu, H. (2017). Deep seismic structure of the northeastern South China Sea: Origin of a high-velocity layer in
the lower crust. Journal of Geophysical Research: Solid Earth, 122, 2831–2858. https://doi.org/10.1002/2016JB013481
Wang, J., Wu, H., & Zhao, D. (2014). P wave radial anisotropy tomography of the upper mantle beneath the North China Craton. Geochemistry,
Geophysics, Geosystems, 15, 2195–2210. https://doi.org/10.1002/2014GC005279
Wang, X., Zhang, J., Fu, P., & Gao, M. (2015). Deep structures of Yishu fault zone derived from gravity data. Seismology and Geology, 37(3),
731–747.
Wang, Y., Forsyth, D. W., & Savage, B. (2009). Convective upwelling in the mantle beneath the Gulf of California. Nature, 462(7272), 499–501.
https://doi.org/10.1038/nature08552
Wei, W., Xu, J., Zhao, D., & Shi, Y. (2012). East Asia mantle tomography: New insight into plate subduction and intraplate volcanism. Journal of
Asian Earth Sciences, 60, 88–103. https://doi.org/10.1016/j.jseaes.2012.08.001
Xia, B., Thybo, H., & Artemieva, I. M. (2017). Seismic crustal structure of the North China Craton and surrounding area: Synthesis and analysis.
Journal of Geophysical Research: Solid Earth, 122, 5181–5207. https://doi.org/10.1002/2016JB013848
Xiong, X., Gao, R., Zhang, X., Li, Q., & Hou, H. (2011). The Moho depth of North China and Northeast China revealed by seismic detection. Diqiu
Xuebao (Acta Geoscientica Sinica), 32(1), 46–56.

LI ET AL. 3300
Geochemistry, Geophysics, Geosystems 10.1029/2018GC007515

Xu, Y. (2014). Recycled oceanic crust in the source of 90–40 Ma basalts in north and northeast China: Evidence, provenance and significance.
Geochimica et Cosmochimica Acta, 143, 49–67. https://doi.org/10.1016/j.gca.2014.04.045
Xu, Y., Chung, S., Ma, J., & Shi, L. (2004). Contrasting Cenozoic lithospheric evolution and architecture in the western and eastern Sino-Korean
Craton: Constraints from geochemistry of basalts and mantle xenoliths. The Journal of Geology, 112(5), 593–605. https://doi.org/10.1086/
422668
Xu, Z., Zhao, Z., & Zheng, Y. (2012). Slab–mantle interaction for thinning of cratonic lithospheric mantle in North China: Geochemical
evidence from Cenozoic continental basalts in central Shandong. Lithos, 146-147, 202–217. https://doi.org/10.1016/j.lithos.2012.05.019
Yao, H., van der Hilst, R. D., & de Hoop, M. V. (2006). Surface-wave array tomography in SE Tibet from ambient seismic noise and two-station
analysis—I. Phase velocity maps. Geophysical Journal International, 166(2), 732–744. https://doi.org/10.1111/j.1365-246X.2006.03028.x
Yao, H., van der Hilst, R. D., & Montagner, J. P. (2010). Heterogeneity and anisotropy of the lithosphere of SE Tibet from surface wave array
tomography. Journal of Geophysical Research, 115, B12307. https://doi.org/10.1029/2009JB007142
Yin, H., Li, J., Zhang, L., Wu, C., & Dong, X. (2008). Analysis of crustal movement features in Shandong area based on the data of GPS
observation network. Northwestern Seismological Journal, 30(3), 276–281.
Zeng, G., Chen, L., Hofmann, A. W., Jiang, S., & Xu, X. (2011). Crust recycling in the sources of two parallel volcanic chains in Shandong, North
China. Earth and Planetary Science Letters, 302(3–4), 359–368. https://doi.org/10.1016/j.epsl.2010.12.026
Zeng, G., Chen, L., Xu, X., Jiang, S., & Hofmann, A. W. (2010). Carbonated mantle sources for Cenozoic intra-plate alkaline basalts in Shandong,
North China. Chemical Geology, 273(1–2), 35–45. https://doi.org/10.1016/j.chemgeo.2010.02.009
Zhang, B., Tang, Y., Xia, T., & Cui, G. (1996). The crustal velocity structure of Liaocheng-Rongcheng—Making an inquiry into “Taishan
earthquake”. Earthquake Research in China, 12(2), 141–146.
Zhang, C., Ma, C., Liao, Q., Zhang, J., & She, Z. (2010). Implications of subduction and subduction zone migration of the Paleo-Pacific plate
beneath eastern North China, based on distribution, geochronology, and geochemistry of late Mesozoic volcanic rocks. International
Journal of Earth Sciences, 100(7), 1665–1684.
Zhao, D. (2004). Origin of the Changbai intraplate volcanism in Northeast China: Evidence from seismic tomography. Chinese Science Bulletin,
49(13), 1401–1408. https://doi.org/10.1360/04wd0125
Zhao, D. (2012). Tomography and dynamics of Western-Pacific subduction zones. Monographs on Environment Earth and Planets, 1(1), 1–70.
https://doi.org/10.5047/meep.2012.00101.0001
Zhao, D., Maruyama, S., & Omori, S. (2007). Mantle dynamics of Western Pacific and East Asia: Insight from seismic tomography and mineral
physics. Gondwana Research, 11(1-2), 120–131. https://doi.org/10.1016/j.gr.2006.06.006
Zhao, D., Mishra, O., & Sanda, R. (2002). Influence of fluids and magma on earthquakes: 571 seismological evidence. Physics of the Earth and
Planetary Interiors, 132(4), 249–267. https://doi.org/10.1016/S0031-9201(02)00082-1
Zhao, D., & Tian, Y. (2013). Changbai intraplate volcanism and deep earthquakes in East Asia: A possible link? Geophysical Journal
International, 195(2), 706–724. https://doi.org/10.1093/gji/ggt289
Zhao, D., Yu, S., & Ohtani, E. (2011). East Asia: Seismotectonics, magmatism and mantle dynamics. Journal of Asian Earth Sciences, 40(3),
689–709. https://doi.org/10.1016/j.jseaes.2010.11.013
Zhao, L., Allen, R. M., Zheng, T., & Zhu, R. (2012). High-resolution body wave tomography models of the upper mantle beneath eastern China
and the adjacent areas. Geochemistry, Geophysics, Geosystems, 13, Q06007. https://doi.org/10.1029/2012GC004119
Zu, J., Wu, Q., & Lian, Y. (1996). The geothermal study of mid-segment of the Tancheng-Lujiang fault zone and its neighboring region.
Earthquake Research in China, 12(1), 43–48.

LI ET AL. 3301

You might also like