measurement of rayleigh wave ellipticity and its

17
Measurement of Rayleigh wave ellipticity and its application to the joint inversion of high-resolution S wave velocity structure beneath northeast China Guoliang Li 1,2 , Haichao Chen 1 , Fenglin Niu 1,3 , Zhen Guo 2,4 , Yingjie Yang 2 , and Jun Xie 2 1 State Key Laboratory of Petroleum Resources and Prospecting, and Unconventional Natural Gas Institute, China University of Petroleum, Beijing, China, 2 Australian Research Council Centre of Excellence for Core to Crust Fluid Systems and GEMOC, and Department Earth and Planetary Science, Macquarie University, North Ryde, New South Wales, Australia, 3 Department of Earth Science, Rice University, Houston, Texas, USA, 4 Institute of Theoretical and Applied Geophysics, Peking University, Beijing, China Abstract We present a new 3-D S wave velocity model of the northeast (NE) China from the joint inversion of the Rayleigh wave ellipticity and phase velocity at 840 s periods. Rayleigh wave ellipticity, or Rayleigh wave Z/H (vertical to horizontal) amplitude ratio, is extracted from both earthquake (1040 s) and ambient noise data (825 s) recorded by the NorthEast China Extended SeiSmic Array with 127 stations. The estimated Z/H ratios from earthquake and ambient noise data show good consistency within the overlapped periods. The observed Z/H ratio shows a good spatial correlation with surface geology and is systematically low within the basins. We jointly invert the measured Z/H ratio and phase velocity dispersion data to obtain a rened 3-D S wave velocity model beneath the NE China. At shallow depth, the 3-D model is featured by strong low-velocity anomalies that are spatially well correlated with the Songliao, Sanjiang, and Erlian basins. The low-velocity anomaly beneath the Songliao basin extends to ~ 23 km deep in the south and ~56 km in the north. At lower crustal depths, we nd a signicant low-velocity anomaly beneath the Great Xingan range that extends to the upper mantle in the south. Overall, the deep structures of the 3-D model are consistent with previous models, but the shallow structures show a much better spatial correlation with tectonic terranes. The difference in sedimentary structure between the southern and northern Songliao basin is likely caused by a mantle upwelling associated with the Pacic subduction. 1. Introduction Although Rayleigh wave dispersion data have been extensively used to study the lithosphere and upper mantle velocity structure [e.g., Shapiro et al., 2005; Yang et al., 2007], it is widely accepted that the relative long-period Rayleigh waves (>8 s) extracted from either earthquakes or ambient noise have very limited resolution on shallow structures. On the other hand, knowledge on shallow velocity structure is of great importance for understanding surface geology [Lin et al., 2012], estimating strong ground motions, as well as better constraining deeper structures with surface data [Waldhauser et al., 2002; Bozdağ and Trampert, 2008; Langston, 2011]. Previous studies have shown that Rayleigh wave ellipticity or Z/H ratio is more sensi- tive to shallow structure than phase velocity dispersion data [e.g., Boore and Toksöz, 1969], allowing us to bet- ter constrain the near-surface shear velocity structure using Z/H ratio. Although the theory of Rayleigh wave ellipticity was well established almost half century ago, its application to study the earth structure seems to fall behind, largely due to lack of high-quality multicomponent seismic data in early days and the difculty in extracting accurate Z/H ratio from real data. Boore and Toksöz [1969] measure the Z/H ratios in frequency domain using earthquake Rayleigh wave data and nd that the observed Z/H ratios are highly scattered. Recently, studies on robust extraction of Z/H ratio and its application to struc- tural inversion [e.g., Tanimoto and Alvizuri, 2006; Ferreira and Woodhouse, 2007; Yano et al., 2009; Lin et al., 2012; Chong et al., 2014; Lin and Schmandt, 2014; Lin et al., 2014] have come under renewed interest because of the rapid increase of high-quality three-component broadband waveform data. Tanimoto and Rivera [2008] propose a cross-correlation-based technique to measure the Z/H ratio in time domain, which they claim to have sufcient accuracy for structural inversion. Lin et al. [2014] nd that short-period Rayleigh wave ellipticities can be robustly extracted from ambient noise cross correlations and construct Z/H ratio maps of western U.S. at periods of 824 s, which exhibits a good correlation with surface geology. LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 1 PUBLICATION S Journal of Geophysical Research: Solid Earth RESEARCH ARTICLE 10.1002/2015JB012459 Key Points: High-resolution S wave velocity model inverted from Rayleigh wave ellipticity and phase velocity The low-velocity anomalies in our model has a better correlation with sedimentary basins Sedimentary depth of Songliao basin associates with a mantle upwelling relate to Pacic subduction Correspondence to: H. Chen, [email protected] Citation: Li, G., H. Chen, F. Niu, Z. Guo, Y. Yang, and J. Xie (2016), Measurement of Rayleigh wave ellipticity and its application to the joint inversion of high-resolution S wave velocity structure beneath northeast China, J. Geophys. Res. Solid Earth, 121, doi:10.1002/2015JB012459. Received 20 AUG 2015 Accepted 6 JAN 2016 Accepted article online 11 JAN 2016 ©2016. American Geophysical Union. All Rights Reserved.

Upload: others

Post on 18-Dec-2021

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Measurement of Rayleigh wave ellipticity and its

Measurement of Rayleigh wave ellipticityand its application to the joint inversionof high-resolution S wave velocitystructure beneath northeast ChinaGuoliang Li1,2, Haichao Chen1, Fenglin Niu1,3, Zhen Guo2,4, Yingjie Yang2, and Jun Xie2

1State Key Laboratory of Petroleum Resources and Prospecting, and Unconventional Natural Gas Institute, China University ofPetroleum, Beijing, China, 2Australian Research Council Centre of Excellence for Core to Crust Fluid Systems and GEMOC, andDepartment Earth and Planetary Science, Macquarie University, North Ryde, New South Wales, Australia, 3Department of EarthScience, Rice University, Houston, Texas, USA, 4Institute of Theoretical and Applied Geophysics, Peking University, Beijing, China

Abstract We present a new 3-D Swave velocity model of the northeast (NE) China from the joint inversionof the Rayleigh wave ellipticity and phase velocity at 8–40 s periods. Rayleigh wave ellipticity, or Rayleighwave Z/H (vertical to horizontal) amplitude ratio, is extracted from both earthquake (10–40 s) and ambientnoise data (8–25 s) recorded by the NorthEast China Extended SeiSmic Array with 127 stations. The estimatedZ/H ratios from earthquake and ambient noise data show good consistency within the overlapped periods.The observed Z/H ratio shows a good spatial correlation with surface geology and is systematically lowwithin the basins. We jointly invert the measured Z/H ratio and phase velocity dispersion data to obtain arefined 3-D S wave velocity model beneath the NE China. At shallow depth, the 3-D model is featured bystrong low-velocity anomalies that are spatially well correlated with the Songliao, Sanjiang, and Erlian basins.The low-velocity anomaly beneath the Songliao basin extends to ~ 2–3 km deep in the south and ~5–6 km inthe north. At lower crustal depths, we find a significant low-velocity anomaly beneath the Great Xing’anrange that extends to the upper mantle in the south. Overall, the deep structures of the 3-D model areconsistent with previous models, but the shallow structures show a much better spatial correlation withtectonic terranes. The difference in sedimentary structure between the southern and northern Songliao basinis likely caused by a mantle upwelling associated with the Pacific subduction.

1. Introduction

Although Rayleigh wave dispersion data have been extensively used to study the lithosphere and uppermantle velocity structure [e.g., Shapiro et al., 2005; Yang et al., 2007], it is widely accepted that the relativelong-period Rayleigh waves (>8 s) extracted from either earthquakes or ambient noise have very limitedresolution on shallow structures. On the other hand, knowledge on shallow velocity structure is of greatimportance for understanding surface geology [Lin et al., 2012], estimating strong ground motions, as wellas better constraining deeper structures with surface data [Waldhauser et al., 2002; Bozdağ and Trampert,2008; Langston, 2011]. Previous studies have shown that Rayleigh wave ellipticity or Z/H ratio is more sensi-tive to shallow structure than phase velocity dispersion data [e.g., Boore and Toksöz, 1969], allowing us to bet-ter constrain the near-surface shear velocity structure using Z/H ratio.

Although the theory of Rayleigh wave ellipticity was well established almost half century ago, its applicationto study the earth structure seems to fall behind, largely due to lack of high-quality multicomponent seismicdata in early days and the difficulty in extracting accurate Z/H ratio from real data. Boore and Toksöz [1969]measure the Z/H ratios in frequency domain using earthquake Rayleigh wave data and find that the observedZ/H ratios are highly scattered. Recently, studies on robust extraction of Z/H ratio and its application to struc-tural inversion [e.g., Tanimoto and Alvizuri, 2006; Ferreira and Woodhouse, 2007; Yano et al., 2009; Lin et al.,2012; Chong et al., 2014; Lin and Schmandt, 2014; Lin et al., 2014] have come under renewed interest becauseof the rapid increase of high-quality three-component broadband waveform data. Tanimoto and Rivera[2008] propose a cross-correlation-based technique to measure the Z/H ratio in time domain, which theyclaim to have sufficient accuracy for structural inversion. Lin et al. [2014] find that short-period Rayleigh waveellipticities can be robustly extracted from ambient noise cross correlations and construct Z/H ratio maps ofwestern U.S. at periods of 8–24 s, which exhibits a good correlation with surface geology.

LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 1

PUBLICATIONSJournal of Geophysical Research: Solid Earth

RESEARCH ARTICLE10.1002/2015JB012459

Key Points:• High-resolution Swave velocity modelinverted from Rayleigh wave ellipticityand phase velocity

• The low-velocity anomalies in ourmodel has a better correlation withsedimentary basins

• Sedimentary depth of Songliao basinassociates with a mantle upwellingrelate to Pacific subduction

Correspondence to:H. Chen,[email protected]

Citation:Li, G., H. Chen, F. Niu, Z. Guo, Y. Yang,and J. Xie (2016), Measurement ofRayleigh wave ellipticity and itsapplication to the joint inversion ofhigh-resolution S wave velocitystructure beneath northeast China,J. Geophys. Res. Solid Earth, 121,doi:10.1002/2015JB012459.

Received 20 AUG 2015Accepted 6 JAN 2016Accepted article online 11 JAN 2016

©2016. American Geophysical Union.All Rights Reserved.

Page 2: Measurement of Rayleigh wave ellipticity and its

Northeast (NE) China is a geological complex area consisting of basins and mountain ranges (Figure 1), andtheir formations are thought to be partly associated with the Pacific subduction. Recently, a large-scalebroadband seismic array, the NECESSArray (NorthEast China Extended SeiSmic Array), was installed in thearea by an international team to study deep subduction and lithosphere structure [Tang et al., 2014]. Usingseismic data from the experiment, Tao et al. [2014] find that teleseismic recordings at stations within thebasins are very complicated because of sediment reverberations, making it difficult to constrain the crustaland mantle structures beneath the stations. Guo et al. [2015] jointly invert receiver function and Rayleighwave data for a 3-D S wave model of NE China and find that the shallow structure (<5 km) is poorly con-strained due to the limited resolution of surface wave dispersion data.

In this study, we present a new 3-D Swave velocity model of NE China derived from joint inversion of Rayleighwave ellipticity and phase velocities at 8–40 s periods. We extract Rayleigh wave ellipticities using both earth-quake and ambient noise data from the NECESSArray, with the short- to intermediate-period measurementsfrom earthquakes and short periods from ambient noise cross-correlation data. Combining the measured Z/Hratios and the phase velocity data, we jointly invert them for a 3-D S wave velocity model beneath NE China.Our 3-D model exhibits strong low-velocity anomalies at shallow depth with its lateral distribution being cor-related very well with the geographic locations of the basins in the area, suggesting that the shallow struc-tures are well constrained by the joint inversion.

2. The NECESSArray Data

The NECESSArray is a temporary broadband array deployed between September 2009 and August 2011 byan international collaborative project. It consists of 127 temporary broadband stations with averaged stationspacing of approximately 80 km (Figure 1) and covers most part of NE China. The Songliao basin, about750 km long and 330–370 km wide, is located at the center of the NE China. It is surrounded by the GreatXing’An range to the west, the Changbai Mountain to the southeast, the Zhangguangcai range to the east,and Lesser Xing’An range to the north (Figure 1). In addition to the Songliao basin, there are other threebasins in NE China, the Hailar basin in the northwest, the Erlian basin in the southwest, and the Sanjiang basinin the northeast (Figure 1).

Figure 1. Map of NE China. The triangles represent the 127 NECESSArray stations used in this study. The yellow trianglesindicate the six stations whose Z/H ratios measured from ambient noise and earthquake data are significantly differentfrom each other and thus are not used in the inversion. The yellow lines outline the four basins, the Songliao, Erlian, Hailar,and Sanjiang in the study area. Major mountain ranges in the area, including the Great Xing’an range, the Lesser Xing’anrange, the Zhangguangcai range, and the Changbaishan volcanic mountain, are also marked in the map.

Journal of Geophysical Research: Solid Earth 10.1002/2015JB012459

LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 2

Page 3: Measurement of Rayleigh wave ellipticity and its

2.1. Ambient Noise Data

Our data processing procedures for the daily continuous ambient noise data are generally adapted from Lin et al.[2014], which are summarized as below. For each seismogram, we first remove the instrument response aftersubtracting the mean and linear trend and decimate the data to one sample per second [Bensen et al., 2007].An antialiasing low-pass filter is applied to the data prior to the down sampling. Following Lin et al. [2014], weapply temporal normalization and spectral whitening to all the three components recorded at each station.After the above preprocessing, we finally calculate nine noise cross-correlation functions (NCFs), Z-Z, Z-N, Z-E,N-Z, N-N, N-E, E-Z, E-N, and E-E, for all possible station pairs. Here the first and second letters refer to the virtual

Figure 2. (a) Configuration of bearings of the raypath at the virtual source (azimuth) and the receiver (back azimuth).(b) Schematic diagrams showing Rayleigh wave signals recorded in the Z and R components by a station pair, station 1 andstation 2. (c) Schematic diagrams showing the four cross correlations of ZZ, ZR, RZ, and RR where station 1 is treated as a virtualsource and station 2 is regular receiver.

Journal of Geophysical Research: Solid Earth 10.1002/2015JB012459

LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 3

Page 4: Measurement of Rayleigh wave ellipticity and its

source and receiver stations, respectively. A rotation matrix is used to rotate the nine NCFs to radial (R) and trans-verse (T) components. The rotation matrix is defined based on virtual source and receiver geometry (Figure 2a):

ZZ

ZR

RZ

RR

0BBB@

1CCCA¼

1 0 0 0

0 cosθr sinθr 0

0 0 0 �cosθs0 0 0 0

0 0 0 0

0 0 0 0

0 0 �sinθs 0

�cosθscosθr �cosθssinθr 0 �sinθscosθr

0

0

0

�sinθssinθr

0BBBBBBB@

1CCCCCCCA

ZZZN

ZE

NZ

NN

NE

EZ

EN

EE

0BBBBBBBBBBBBBBBBBB@

1CCCCCCCCCCCCCCCCCCA

(1)

where θs and θr are the bearings of the raypath at the virtual source (azimuth) and the receiver (back azimuth),respectively (Figure 2a). An example of NCFs calculated for the station pair, NE12-NE46, is shown in Figures 3aand 3b, which, respectively, show the positive and negative sides of the empirical Green functions filtered withan 8–25 s band-pass filter. The corresponding particle motions are shown in Figures 3c and 3d.

Figures 3c and 3d show the particle motions of the Rayleigh wave at station NE46, excited by virtual verticaland horizontal sources, respectively. Similarly, Figures 3e and 3f represent the particle motions of Rayleighwaves recorded at station NE12. Both stations exhibit an elliptical particle motion; however, the basin stationNE46 (Figure 1) is characterized by a horizontal major axis, while the Great Xing’an range station NE12(Figure 1) possesses a vertical major axis, indicative of significant difference in shallow structure beneaththe two stations.

2.2. Earthquake Data

To measure Z/H ratio from earthquake data, we select ~400 earthquakes occurring between September 2009and August 2011 with epicentral distances ranging between 10 and 140° and magnitudes greater than 6.0.We cut out the Rayleigh wave of each earthquake, remove the linear trend and mean for each component,and then band pass filtered the data at 8–50 s periods. We further decimate the data to one sample persecond after applying an antialiasing filter to the data and rotate the two horizontal components to the radialand transverse directions. The radial and vertical components are finally selected to measure Z/H ratio in theperiod range of 10–40 s, which partially overlaps with the ambient noise period band of 8–25 s.

3. Methodology3.1. Z/H Ratio Measurement

For the NCFs calculated for one virtual source-receiver station pair, the positive part is associated with theGreen’s function traveling from the virtual source to the receiver, and therefore, the Z/H ratios measured fromthis part of the NCFs reflect the velocity structure beneath the receiver station. Similarly, the Z/H ratiosmeasured from the negative side of the NCFs are expected to be associated with the velocity structurebeneath the virtual source station. Thus, we did not stack the positive and negative sides of the NCFs butstacked the Green’s functions from a vertical and radial source to enhance the signal-to-noise ratio (SNR)in measuring the Z/H ratio (Appendix A).

The ZZ and ZR NCFs can be considered as the vertical and radial records at the receiver station due to a verticalforce at the virtual source. Since we apply the normalization and whitening to the noise data of all the threecomponents with the same temporal normalization functions and spectrum, the relative amplitude of thetwo NCFs is preserved in the cross-correlation calculation. Similarly, the RZ and RR NCFs are the vertical andradial records of radially directed force. Theoretically, Rayleigh wave radiated from a vertical or a radial pointforce is expected to have the same Z/H ratio. Since the maximum amplitude of the four NCFs (ZZ, ZR, RZ,

Journal of Geophysical Research: Solid Earth 10.1002/2015JB012459

LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 4

Page 5: Measurement of Rayleigh wave ellipticity and its

and RR) is in the same order of magnitude, therefore, we use the sum of ZZ and RZ as the numerator and thesum of ZR and RR as the denominator in computing the Z/H ratio:

E Tð Þ ¼ AZZAZR

¼ ARZARR

¼ A H ZZ½ �þRZð ÞA H ZZ½ �þRRð Þ

; (2)

Figure 3. Examples of the calculated NCFs of ZZ, ZR, RZ, and RR between station pair NE12 and NE46 with (a) NE12 as avirtual source and NE46 as a receiver and (b) vice versa. The NCFs are band-pass filtered at 8–25 s periods. (c and d)Particle motions of the Rayleigh wave recorded at NE46, excited by a vertical and radial source, respectively. (e and f) Sameas Figures 3c and 3d except for NE12. Note that the low Z/H ration station NE12 and the high Z/H station NE46 are located atthe Great Xing’an range and the Songliao basin, respectively.

Journal of Geophysical Research: Solid Earth 10.1002/2015JB012459

LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 5

Page 6: Measurement of Rayleigh wave ellipticity and its

where E(T) is the Z/H ratio for period T, and A indicates the maximum amplitude of the Rayleigh wave. H[ZZ]and H[ZR] represent the Hilbert transform of the ZZ and ZR NCFs.

It is known that there is a 90° phase shift between the Z and R components (R minus Z) for Rayleigh wavesrecorded at both the virtual source and receiver stations (Figure 2b). Their cross-correlation functions, ZZand RR NCFs, are expected to be in phase, while the ZR and RZ NCFs are, respectively, 90° ahead and 90°behind the ZZ and RR NCFs, i.e., for a monotonic Rayleigh wave with a period T, the arrival times at the fourNCFs are expected to have the following relationship (Figure 2c):

tZZ ¼ tRR ¼ tRZ � T=4 ¼ tZR þ T=4 (3)

Therefore, in order to stack ZZ and RZ, a Hilbert transform is applied to ZZ to keep the two NCFs in phase.Similarly, we apply a Hilbert transform to ZR to delay its phase by 90° to align it with RR and then sum thetwo as the denominator.

For the selected Rayleigh wave time window, we first employ a phase-matched filter [Levshin and Ritzwoller,2001] to the stacked Z and R components to eliminate noises from the data. The algorithm for estimating theZ/H ratio is adapted from Tanimoto and Rivera [2008]. Here we briefly describe the major steps. We first filterthem with a series of Gaussian band pass with bandwidths that vary from 0.0025 to 0.0400Hz depending onthe central frequencies and source-receiver distances [Herrmann, 1973]. For each band-pass-filtered data, wecompute the zero-lag cross-correlation coefficient of the vertical component and the phase-delayed radialcomponent. We discard the data that have a zero-lag cross-correlation coefficient less than 0.8. If the narrowlyband-pass-filtered data are selected, the envelopes of the two components are computed, and their maximaare used to calculate the Z/H ratio. Since we use the maximum of each component in computing the Z/Hratios, the phase-matched filtering appears to have little effect on our measurements.

To retain the most reliable measurements, we adopt the following additional criteria to further select the data.

1. We only select Rayleigh wave data with a SNR ≥ 8 on both the stacked Z and R components. Here the SNRis defined as the max amplitude of the signal to the RMS of a 500 s long noise window defined from 3000 safter the Rayleigh wave [Bensen et al., 2007].

2. We only select station pairs with an interstation distance greater than three wavelengths to satisfy the far-field requirement for the ambient noise cross-correlation data [Bensen et al., 2007].

3. For each period, we require at least 20 Z/H measurements in order to get statistically robust average.Furthermore, we require that the distribution of the Z/H measurements roughly follow a Gaussian distri-bution with a standard error less than 15% of the average.

4. We further discard stations with less than five robust average Z/H estimates (i.e., at five periods).

We finally select 110 stations that satisfy the above criteria (i.e., 17 stations are dropped off), among which sixstations show significant differences (≥10%) in the measurement Z/H ratio between the NCFs and earthquakeevents data. We have examined possible causes for the observed discrepancy, including sensor misorienta-tion and bias from event azimuthal distribution, but fail to obtain a good explanation. Therefore, we decide toleave this issue for further investigation and exclude these six stations in the joint inversion.

3.2. Joint Inversion of Dispersion and Z/H Ratio Data

We jointly invert phase velocity dispersion curves and ellipticities at period 8–40 s using a DRAM (DelayedRejection Adaptive Metropolis) algorithm [Haario et al., 2006; Mira, 2001] based on Markov chain MonteCarlo (MCMC) method [Bodin et al., 2012; Afonso et al., 2013; Shen et al., 2013] to obtain 1-D S wave velocitydepth profiles for each station. The Bayesian MCMC inversion method has been well discussed in several lit-eratures, such as Bodin et al. [2012], Afonso et al. [2013], and Shen et al. [2013] and thus will not be describedhere. The basic principle of DRAM algorithm is to improve the MCMC algorithm by updating the acceptanceprobability of the new candidate model on a regular interval determined by previous samples in the chain,which could increase the probability of sampling regions with high posterior probability.

The phase velocity dispersion data are obtained from Guo et al. [2015]. Phase velocities at different period werefirst measured from ambient noise data and then inverted to 0.2° by 0.2° spatial grids across NE China. Thedispersion data at each station are taken from the phase velocities of the closest grid. The Z/H ratios are mea-sured at each station and are obtained from ambient noise data for periods less than 25 s and from earthquakeevents for periods longer than 25 s. Both data sets are expected to reflect mainly the seismic velocity structurebeneath each station and thus can be jointly inverted to obtain the seismic structure underneath the stations.

Journal of Geophysical Research: Solid Earth 10.1002/2015JB012459

LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 6

Page 7: Measurement of Rayleigh wave ellipticity and its

The 1-D shear velocity model consists of a sedimentary layer, a crystalline crustal layer, and a mantle layeralong the depth direction, which is delineated by 13 independent parameters (Table 1). The sediment layerhas three parameters: thickness and velocities at top and bottom of the layer. The crustal layer is described byfive parameters including Moho depth and four cubic B-splines to represent the vertical variation of Vs. Themantle is characterized by five cubic B-splines to represent Vs variations. For the sedimentary layer, we set Vp/Vs ratio to 2.0 and initial sediment thickness is taken from Laske and Masters [1997]. The crustal Vp/Vs ratio andinitial Moho depth are taken from Tao et al. [2014]. In the upper mantle, we choose a constant Vp/Vs ratio of1.732. We also employ Birch’s law of Vp= 3.125ρ� 2.40 (ρ: g/cm3, Vp, Vs: km/s) [Birch, 1961] to tie density ρ toVs to be inverted through Vp. A prior constraint is imposed to the Vs contrasts across the sediment base andthe Moho, forcing them to be positive (i.e., sedimentary Vs<upper crust Vs and lower crust Vs<mantle Vs).Phase velocity dispersion curves and ellipticities are modeled using the MINEOS [Masters et al., 2007]. Themisfit of the joint inversion (MJOINT) is defined as

:

MJOINT ¼ c1�MSW þ c2�MZH

¼ c1�XNi¼1

Gi mð Þ � Dobsi

� �2σ2i

þ c2�XMj¼1

Rj mð Þ � Bobsj

h i2

γ2j

(4)

Here Gi(m) is the calculated phase velocity frommodelm at period i on a discrete grid of N periods. Diobs is the

observed phase velocity at the same period. The Rj(m) is the predicted Z/H ratio of model m at period j on adiscrete grid ofM periods, and Bj

obs is the observed Z/H ratio at the same period. σi and γj represent the uncer-tainties of phase velocity at period i and Z/H ratio at period j, respectively. The weighting factors of two datasets are set to be equal (c1 = c2 = 0.5).

The initial models are constructed based on the results of Guo et al. [2015] for the upper 80 km and the pre-liminary reference Earth model [Dziewonski and Anderson, 1981] between the depths of 80 and 200 km. Below200 km, the velocities are set to be constant. We also assume an isotropic velocity model in the joint inver-sion. Since the longest period of our data is 40 s (Figure 8b), which has a depth resolution limited to theuppermost mantle, thus we only invert Vs above 80 km.

For each station, we first randomly generate 100,000models and compute their misfits according to equation(4). We then choose 2000 models with the lowest misfits to estimate the final inversion result. The final 1-Dvelocity is obtained by averaging the 2000 models.

4. Results and Discussion4.1. The Measured Z/H Ratios

Figure 4 shows the histogram of Z/H ratio of 15 s Rayleigh wave measured from NCFs (Figure 4a) and earth-quake data (Figure 4b) at station NE11, with both data sets following the Gaussian distribution. We computetheir averages and standard deviations of the means and use them as the Z/H ratio estimatesand uncertainties.

Figure 5 shows the Z/H ratios at six stations measured from NCFs (red solid points) and earthquake data (bluepoints) with the estimated error bars. Stations NE11, NE46, and NE73 are located inside the basins, whilestations NE12 and NE53 are based on the Great Xing’an range. The measured Z/H ratios at the basin stationsappear to be relatively low at short periods (10–20 s) but increase substantially with increasing period. On the con-trary, the Z/H ratios measured from mountainous areas are comparatively higher at short periods (10–20 s) but

Table 1. Model Space and References

Model Parameters Range Reference

Sediment thickness 0–2m0 (km) Laske and Masters [1997]VSV, top of sedimentary layer 1.0–3.5 (km) Guo et al. [2015]VSV, bottom of sedimentary layer 1.0–3.5 (km) Guo et al. [2015]Crystalline crustal thickness m0 ± 5 (km) Tao et al. [2014]B-spline coefficients, crust m0 ± 0.2m0 (km/s) Guo et al. [2015]B-spline coefficients, mantle m0 ± 0.2m0 (km/s) Dziewonski and Anderson [1981]

Journal of Geophysical Research: Solid Earth 10.1002/2015JB012459

LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 7

Page 8: Measurement of Rayleigh wave ellipticity and its

(a) (b)

Figure 4. The distributions of all Z/H ratios of 15 s Rayleigh wave measured from (a) ambient noise and (b) earthquake dataat station NE11. The calculated average value and standard error are listed at the top right corner of each plot.

Figure 5. The measured Z/H ratios at six stations (see their location in Figure 1) as a function of period, with blue and redpoints representing measurements from ambient noise and earthquake data, respectively. Uncertainties are indicated byvertical bars. These stations are (a) NE11, (b) NE12, (c) NE46, which are located in the southern Songliao basin, and (d) NE53,(e) NE73, (f) NE34 that are situated at the Great Xing’an range. Note the Z/H ratios at station NE34 in Figure 5f measured fromNCFs and earthquake data differ significantly from each other.

Journal of Geophysical Research: Solid Earth 10.1002/2015JB012459

LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 8

Page 9: Measurement of Rayleigh wave ellipticity and its

decrease slightly with increasing period. At longer periods (>30 s), the Z/H ratios become relatively stable in boththe mountain and basin areas. One interesting phenomenon is that the Z/H ratios measured from NCFs seem tobe smoother than those from earthquake data at short periods, possibly related to the fact that short-periodRayleigh waves from earthquakes are noisier resulting from scattering and attenuation after long-distance

Figure 6. Maps of the Z/H ratios at four different periods, 10, 15, 20, and 25 s, measured from (a, c, e, and g) ambient noiseand (b, d, f, and h) earthquake data, respectively.

Journal of Geophysical Research: Solid Earth 10.1002/2015JB012459

LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 9

Page 10: Measurement of Rayleigh wave ellipticity and its

propagation. In general, the Z/H ratios measured from NCFs and earthquake data agree well with each other atmost stations, except a few stations that are marked by yellow triangles in Figure 1. Figure 5f shows one exampleof such station, NE34, whose Z/H ratios measured from the two types of data sets are significantly different. Asmentioned before, we have investigated the causes for such a discrepancy and are not able to find an obviousissue to explain it; thus, we would like to leave this to future studies.

In Figure 6, we present maps of the observed of Z/H ratios at four periods measured by ambient noise andearthquake data. In each map, triangles denote the stations at which the measured Z/H ratios are used forthe interpolation. The left panel shows the maps of the Z/H ratios measured from NCFs, while the right col-umn are the corresponding Z/H ratios estimated from earthquake data. A comparison of the two columnssuggests that the Z/H ratios derived from earthquake data (Figures 6a, 6c, 6e, and 6g) are generally higherthan those estimated from NCFs (Figures 6b, 6d, 6f, and 6h). One possible reason could be related to off greatcircle path propagation effects caused by structural anomalies, which are expected to be larger in the earth-quake data due to the relatively longer raypaths. The off great circle path propagation can consistently lowerthe radial amplitude and therefore results in higher than actual Z/H ratios. However, the difference betweenthe NCFs and earthquake data are small that the two columns seem to be very similar to each other, suggest-ing that Z/H ratios can be robustly extracted from the two types of data with proper processing techniques.All the maps show that the Z/H ratios inside the Songliao basin are distinctly lower as compared to itssurrounding areas. This contrast fades away gradually with increasing period.

We adopt the method proposed by Lin et al. [2009] to estimate the uncertainties of Z/H ratio. Lin et al. [2009]suggest the uncertainty in the measured Z/H ratio can be approximated by 1.5 times of the standard devia-tion from the statistical mean. Figure 7 presents the uncertainty maps of the Z/H ratios computed with theabove method at periods 10 and 25 s using measurements from the NCFs and earthquake data. At 10 s,the measurement with earthquake data exhibits much larger uncertainties, especially inside the basins, as

Figure 7. Maps showing the uncertainties of the Z/H ratio measurement at periods (a and b) 10 s and (c and d) 25 s, usingNCFs (Figures 7a and 7c) and earthquake data (Figures 7b and 7d), respectively.

Journal of Geophysical Research: Solid Earth 10.1002/2015JB012459

LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 10

Page 11: Measurement of Rayleigh wave ellipticity and its

compared to the ambient noise data (Figures 7a versus 7b). But at 25 s, the uncertainties in the Z/H ratiomeasurements from the two types of data are comparable and smaller.

4.2. The 3-D Velocity Model and Discussions

Figure 8 shows the sensitive kernels of the phase and ellipticity at period 8 s and 40 s. Those kernels are pro-duced during the joint inversion of station NE11. As mentioned above, we employ a fixed Vp/Vs ratio(Vp=2.0Vs for the sedimentary layer or Vp= 1.732Vs for other layers) and a scaling law between velocityand density (ρ= (2.400 + 1.732Vs)/3.125 for sedimentary layer or ρ= (2.400 + 1.732Vs)/3.125 for other layers)in our inversion for S wave velocity beneath each station. Here we also use these assumed scaling relation-ships in computing the phase velocity and Z/H ratio kernels. Our main purpose here is to illustrate the sensi-tivity difference to depth between the two types of kernels; thus, sensitivity kernels to the Pwave velocity anddensity are not shown here. The two phase velocity kernels reach their maxima at roughly one third of theirwavelengths. The two Z/H kernels, on the other hand, have their peak sensitivity at zero depth, whichexplains why Z/H ratios are better observables to constrain shallow structures than the dispersion data.

At each station, we jointly invert the dispersion and Z/H ratio data for a 1-D S wave velocity model. Figure 9shows two examples of the inverted 1-D velocity models beneath station NE12 and NE46. Both the dispersionand Z/H ratio data are well fitted. Once we invert the 1-D velocity models at all the stations, we employ amini-mum curvature surface-fitting method [Smith and Wessel, 1990] to interpolate the results onto 0.5° by 0.5°grids covering our study region from 41° to 48.5° in latitude and from 115° to 134° in longitude to formthe final 3-D velocity model. We remove areas that are more than 150 km away from the nearest station.

Figure 10a shows the velocity variations at 1 km beneath the surface. For comparison, we also show the Swave velocity at the same depth inverted from phase dispersion data alone in Figure 10b. We notice two dis-tinct differences between the two maps. The S wave velocities inside the basins derived from the joint inver-sion are significantly lower than those from the dispersion inversion. The lowest Swave velocities at the 1 kmdepth obtained by the joint and dispersion inversions are, respectively, 1.22 and 1.72 km/s, and the averagedS wave velocities within the Songliao basin at the same depth are 1.91 and 2.36 km/s. The differences are atthe ~20–35% level. The boundaries that outline the low-velocity anomalies are very sharp and match verywell with the basins, such as the Songliao, Sanjiang, and Erlian basins. There is only one station located inside

(a) (b)

Figure 8. Sensitive kernels of phase velocity (red line) and Z/H ratio (blue line) at periods of 8 s and 40 s shown in the(a) left and (b) right, respectively. Note the difference in sensitive depth between these two types of data.

Journal of Geophysical Research: Solid Earth 10.1002/2015JB012459

LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 11

Page 12: Measurement of Rayleigh wave ellipticity and its

the Hailar basin; thus, the structure beneath it is not well resolved in this study. Such an excellent correlationbetween the low-velocity areas and the basin, however, is not so clear in the velocity map (Figure 10b), whichis inverted from phase velocity dispersion data only. The comparison clearly demonstrates the resolvingpower of the Z/H ratio data on shallow structures.

This excellent correlation between the Songliao basin and the strong low-velocity anomaly extends to at least4 km deep, where we observe a NW-SE trending ridge with normal velocities that separates the basin into thenorthern and southern parts (Figure 10c). The slow anomaly beneath the northern Songliao basin appears toextend to at least ~6–7 km (Figures 10d and 10e). The difference in sedimentation between the northern andsouthern Songliao basin can be further seen in the depth cross sections plotted in Figure 11. The geographiclocations of these cross sections are shown in Figure 10c. AA′ is an EW section taken along the 43.5°N latitude,which samples the southern Songliao basin. The low-velocity sediment is approximately 3 km thick in thewest side and shallows gradually toward east, where the sediment is ~ 2 km thick (Figure 11a). The BB′ linesampling the northern Songliao basin shows a similar westward dipping trend (Figure 11b). The sedimentsin the west are, however, twice as thick as those in the southern Songliao basin. The S wave velocities at shal-low depth beneath the northern Songliao basin (~1.4 km/s) are nearly 30% lower than those in the southernSongliao basin (1.8 km/s). We notice that these two depth sections agree very well with the results of twoactive seismic profiles roughly at the same locations shown in Figure 3 of Wei et al. [2010]. The contrastbetween the southern and northern Songliao basin can be further seen in the CC′ section (Figure 11c), whichsamples the basin along its NE-SW strike. In Figure 11d, we further show a NEE-SWW trending depth sectionthat samples the Erlian, northern Songliao, and Sanjiang basins. The low-velocity sediments are clearly shownbeneath the three basins.

Figure 9. Two examples of the joint inversion combining dispersion and Z/H ratio data. (a and b) The fitting of Z/H ratio and phase velocities at station NE12. Blackpoints with vertical error bars connected by thin black lines represent the measured data, while the red lines indicate Z/H ratios and phase velocities calculated fromthe inverted S wave velocity model. (c and d) Similar to Figures 9a and 9b except for station NE46. (e) The final 1-D S wave velocity models beneath the two stations,NE12 and NE46, are shown in red and green, respectively. The two models are calculated by averaging the velocity structure of the 2000 best fitting models.

Journal of Geophysical Research: Solid Earth 10.1002/2015JB012459

LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 12

Page 13: Measurement of Rayleigh wave ellipticity and its

In the lower crustal depth, the most distinct feature is the low velocity beneath the great Xing’an range(Figure 10f). This low-velocity anomaly becomes even stronger at 32 km deep (Figure 10 g). At this depth,we also observe a very strong high-velocity structure at the eastern edge of the Songliao basin where thethinnest crust of the study area is located based on Tao et al. [2014]. Thus, the velocity contrast betweenthe eastern Songliao and great Xing’an reflects the velocity difference between mantle and crust. In theuppermost mantle (Figures 10 h and 10i), low-velocity anomalies beneath the southern Xing’an range, theJiamusi Massif, and the Changbaishan volcanic complex are the dominant features, which are clearly linkedto magmatic activities.

The most distinct features shown in the jointly invert 3-D velocity model are the substantial low velocitybeneath the basins and the strong contrast in sedimentary structure between the southern and northernSongliao basin. The southern Songliao basin is generally considered as a wide rift system developed fromthe Late Jurassic to Early Cretaceous [Wei et al., 2010]. Previous geologic and active source seismic studies[e.g., Feng et al., 2010; Wei et al., 2010] suggest that the southern Songliao basin is marked by substantialsynrift subsidence but insignificant postrift subsidence. The northern Songliao basin, on the other hands,appears to have a distinct tectonic subsidence history characterized by significant synrift and postriftsedimentation; the latter can reach as thick as 4 km. The insignificance of postrift subsidence of the southern

Figure 10. Inverted 3D S wave model plotted at various depths. (a and b) S wave velocities at 1 km deep inverted from phase velocity and Z/H ratio data together andphase velocity data alone are shown for comparison. Note the excellent correlation between the low-velocity structures and the basins in the area in Figure 10a. (c–i)Depth slices of the jointly inverted 3-D Swave velocitymodel at 4, 6, 7, 24, 32, 40, and 50 km. The depth of each slice is labeled at the lower right corner of each figure. Redtriangles are stations whose 1-D S wave velocity models are used to synthesize the final 3-D model. The basins are outlined by dark grey lines. EL, HL, SL, and SJ inFigure 10b represent the Erlian, Hailar, Songliao, and Sanjiang basins, respectively. The four dashed lines in Figure 10c indicate the geographic locations of the four depthcross sections shown in Figure 11.

Journal of Geophysical Research: Solid Earth 10.1002/2015JB012459

LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 13

Page 14: Measurement of Rayleigh wave ellipticity and its

Figure 11. Four cross-sections of the jointly inverted 3-D S wave model. The geographic locations of these cross sectionsare delineated in Figure 10c. (a and b) Swave velocities across the southern and northern Songliao basin. (c) A cross sectionthat samples the Songliao Basin along its strike direction, illustrating the difference in sedimentary structure betweensouthern and northern Songliao basin. (d) A NEE-SWW trending depth section that shows the structure beneath the Erlianand northern Songliao and Sanjiang basins.

Journal of Geophysical Research: Solid Earth 10.1002/2015JB012459

LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 14

Page 15: Measurement of Rayleigh wave ellipticity and its

Songliao basin is difficult to invoke pure shear extensional mechanism [McKenzie, 1978], which predicts aconsiderable thermal subsidence following initial rifting. Wei et al. [2010] proposed that the low subsidencerate observed in the southern Songliao basin could be cause by inflow of low-crustal materials fromsurrounding highly elevated mountain regions. The great Xing’an range is characterized by large-scaleMesozoic magmatism, which is dominantly granitic as observed at the surface. Granitic rocks with severalkilometers thick cover most part of the Great Xing’an range [Wu et al., 2003a, 2003b]. In particular, we observea strong low-velocity anomaly beneath the southern end of the Great Xing’an range, which extends from themiddle crust to the uppermost mantle. This suggests that temperature at lower crustal depths is relativelyhigh, which can significantly reduce the viscosity of lower crust, making it easy to flow to its easternneighbors, the southern Songliao basin that has a low elevation. Thus, our 3-D velocity model is consistentwith this lower crustal flow mechanism to explain the low level of postrift sedimentation inside the southernSongliao basin. However, Tao et al. [2014] argued that this scenario has a problem in explaining the low tonormal crustal Vp/Vs ratio in the southern Songliao basin. In general, if crustal thickening were accomplishedby injection of lower crustal materials, then one would expect to observe a positive correlation betweencrustal thickness and Vp/Vs ratio, as mafic lower crustal rocks tend to have high Vp/Vs ratios. The positivecorrelation was, however, not observed in the measured Vp/Vs ratio.

The small postrift subsidence can be due to uplift forces coming from the mantle. Tao et al. [2014] found thatsouthern Songliao basin has a slightly positive residual topography (~0.25 km) that cannot be explained bythe Airy isostasy model. Volcanic rocks erupted during 86–39 aremainly found in the southern Songliao basin[Liu et al., 2001], suggesting that a mantle upwelling was present beneath that part of the basin after therifting in late Jurassic and early Cretaceous, which jeopardized the afterward thermal subsidence that haveotherwise occurred. Liu et al. [2015] found significant difference in the subduction geometry of the Pacificslab in the transition zone between the southern and northern Songliao basin. The Pacific slab is understagnant subduction in the north but not in the south. Slab segmentation might have occurred beneaththe southern Songliao basin, allowing hot asthenospheric material trapped under the subducting Pacificplate to rise to the upper mantle through the edge of the Pacific slab near the upper and lower mantleboundary [Tang et al., 2014]. Thus, we believe that the low postrift subsidence rate in the southernSongliao basin has a plausible deep origin in the mantle.

5. Conclusions

In this study, we measure the Rayleigh wave Z/H ratio in the period band of 8–40 s from ambient noiseand earthquake data recorded by the NECESSArray in NE China. At short periods, measured Z/H ratio atbasin stations is significantly lower than those in surrounding areas. We jointly invert the measured Z/Hratios and phase velocity dispersion data for a high-resolution 3-D S wave velocity model of NE China.Compared to the 3-D model inverted from dispersion data alone, our model exhibits the followingtwo distinct features: (1) the shallow low-velocity anomalies possess much stronger amplitude and havea better correlation with the sedimentary basins in the area; (2) the southern Songliao basin is charac-terized by thin and moderately low-velocity sedimentary structure, while the sediments in the northernSongliao basin are thick and of very lower S wave velocity, a striking contrast that is consistent with pre-vious results derived from active source P wave data. Combining the results from previous passivesource seismic studies [e.g., Tao et al., 2014; Liu et al., 2015], we propose that the insignificant amountof postrift subsidence observed inside the southern Songliao basin is likely caused by a mantle upwel-ling associated with the Pacific subduction.

Appendix A

For a station pair A and B, we first use station B as the reference station to compute the cross correlations:ZBZA, ZBRA, RBRA, and RBRA. Here A and B represent stations and Z and R indicate the vertical and radial direc-tions. The positive parts of the four cross correlograms (ZBZA

+, ZBRA+, RBRA

+, and RBRA+) can be considered as

the vertical and radial components of the Rayleigh waves recorded at station A generated by a point sourceat station B in the vertical and radial directions, respectively. Based on reciprocity principle of Green’sfunction, their negative sections (ZBZA

�, RBZA�, ZBRA

�, and RBRA�) can be considered as the vertical and radial

Journal of Geophysical Research: Solid Earth 10.1002/2015JB012459

LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 15

Page 16: Measurement of Rayleigh wave ellipticity and its

components of the Rayleigh waves recorded at station B generated by a point source at station A in the ver-tical and radial directions, i.e.,

ZBZA� ¼ ZAZB

þ

RBZA� ¼ ZARB

þ

ZBRA� ¼ RAZB

þ

RBRA� ¼ RARB

þ:

(A1)

This is also true for the cross correlograms computed with station A as the reference station, ZAZB, ZARB, RARB,and RARB, for which the negative portions can be considered as the Green’s functions at station A from asource at B:

ZAZB� ¼ ZBZA

þ

RAZB� ¼ ZBRAþ

ZARB� ¼ RBZAþ

RARB� ¼ RBRAþ:

(A2)

Therefore, the Z/H ratio at station A can be computed using either the two positive parts of the BAcorrelograms

EA ¼ AZBZAþ

AZBRAþ¼ ARBZAþ

ARBRAþ; (A3a)

or the two negative parts of the AB correlograms:

EA ¼ AZAZB�

ARAZB�¼ AZARB�

ARARB�: (A3b)

Based on equation (A2), equation (A3a) is identical to equation (A3b); thus, the positive parts of the BAcorrelograms are enough to estimate the Z/H ratio at station A.

ReferencesAfonso, J. C., J. Fullea, W. L. Griffin, Y. Yang, A. G. Jones, J. A. D. Connolly, and S. Y. O’Reilly (2013), 3-D multiobservable probabilistic inversion

for the compositional and thermal structure of the lithosphere and upper mantle. I: A priori petrological information and geophysicalobservables, J. Geophys. Res. Solid Earth, 118, 2586–2617, doi:10.1002/jgrb.50124.

Bensen, G. D., M. H. Ritzwoller, M. P. Barmin, A. L. Levshin, F. Lin, M. P. Moschetti, M. P. Moschetti, N. M. Shapiro, and Y. Yang (2007), Processingseismic ambient noise data to obtain reliable broad-band surface wave dispersion measurements, Geophys. J. Int., 169(3), 1239–1260,doi:10.1111/j.1365-246X.2007.03374.x.

Birch, F. (1961), The velocity of compressional waves in rocks to 10 kilobars: 2, J. Geophys. Res., 66(7), 2199–2224, doi:10.1029/JZ066i007p02199.

Bodin, T., M. Sambridge, H. Tkalčić, P. Arroucau, K. Gallagher, and N. Rawlinson (2012), Transdimensional inversion of receiver functions andsurface wave dispersion, J. Geophys. Res., 117, B02301, doi:10.1029/2011JB008560.

Boore, D. M., and M. N. Toksöz (1969), Rayleigh wave particle motion and crustal structure, Bull. Seismol. Soc. Am., 59(1), 331–346.Bozdağ, E., and J. Trampert (2008), On crustal corrections in surface wave tomography, Geophys. J. Int., 172(3), 1066–1082, doi:10.1111/

j.1365-246X.2007.03690.x.Chong, J., S. Ni, and L. Zhao (2014), Joint inversion of crustal structure with the Rayleigh wave phase velocity dispersion and the ZH ratio,

Pure Appl. Geophys., doi:10.1007/s00024-014-0902-z.Dziewonski, A. M., and D. L. Anderson (1981), Preliminary reference Earth model, Phys. Earth Planet. Inter., 25(4), 297–356, doi:10.1016/

0031-9201(81)90046-7.Ferreira, A. M. G., and J. H. Woodhouse (2007), Observations of long period Rayleigh wave ellipticity, Geophys. J. Int., 169, 161–169,

doi:10.1111/j.1365-246X.2006.03276.x.Feng, Z., C. Jia, X. Xie, S. Zhang, Z. Feng, and T. A. Cross (2010), Tectonostratigraphic units and stratigraphic sequences of the nonmarine

Songliao basin, northeast China, Basin Res., 22(1), 79–95, doi:10.1111/j.1365-2117.2009.00445.x.Guo, Z., Y. J. Chen, J. Ning, Y. Feng, S. P. Grand, F. Niu, H. Kawakatsu, S. Tanaka, M. Obayashi, and J. Ni (2015), High resolution 3-D crustal

structure beneath NE China from joint inversion of ambient noise and receiver functions using NECESSArray data, Earth Planet. Sci. Lett.,416, 1–11, doi:10.1016/j.epsl.2015.01.044.

Haario, H., M. Laine, A. Mira, and E. Saksman (2006), DRAM: Efficient adaptive MCMC, Stat. Comput., 16(4), 339–354.Herrmann, R. B. (1973), Some aspects of band-pass filtering of surface waves, Bull. Seismol. Soc. Am., 663(2), 663–671.Laske, G., and G. Masters (1997), A global digital map of sediment thickness, Eos Trans. AGU, 78, 483.Langston, C. A. (2011), Wavefield continuation and decomposition for passive seismic imaging under deep unconsolidated sediments, Bull.

Seismol. Soc. Am., 101(5), 2176–2190, doi:10.1785/0120100299.Levshin, A. L., and M. H. Ritzwoller (2001), Automated detection, extraction, and measurement of regional surface waves, in Monitoring the

Comprehensive Nuclear-Test-Ban Treaty: Surface Waves, edited by A. L. Levshin and M. H. Ritzwoller, pp. 1531–1545, Birkhäuser, Basel.Lin, F.-C., and B. Schmandt (2014), Upper crustal azimuthal anisotropy across the contiguous U.S. determined by Rayleigh wave ellipticity,

Geophys. Res. Lett., 41, 8301–8307, doi:10.1002/2014GL062362.

Journal of Geophysical Research: Solid Earth 10.1002/2015JB012459

LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 16

AcknowledgmentsWe would like to thank all the peopleinvolved in the NECESSArray projectand Steve Grand from UT Austin fortheir helpful discussion. The data of theNECESSArray project used by this paperare available at http://ds.iris.edu. Wealso thank Fan-Chi Lin and anotheranonymous reviewer for their construc-tive comments and suggestions, whichsignificantly improved the quality of thispaper. G.L., H.C., F.N. are supported byNSFC 41274099, Z.G., Y.Y., and J.X. aresupported by the Australian ResearchCouncil Discovery grant (DP120103673).This is contribution 695 from the ARCCentre of Excellence for Core to CrustFluid Systems (www.ccfs.mq.edu.au)and 1052 in the GEMOC Key Centre(www.gemoc.mq.edu.au).

Page 17: Measurement of Rayleigh wave ellipticity and its

Lin, F.-C., M. H. Ritzwoller, and R. Snieder (2009), Eikonal tomography: Surface wave tomography by phase front tracking across a regionalbroadband seismic array, Geophys. J. Int., 177(3), 1091–1110, doi:10.1111/j.1365-246X.2009.04105.x.

Lin, F.-C., B. Schmandt, and V. C. Tsai (2012), Joint inversion of Rayleigh wave phase velocity and ellipticity using USArray: Constrainingvelocity and density structure in the upper crust, Geophys. Res. Lett., 39, L12303, doi:10.1029/2012GL052196.

Lin, F.-C., V. C. Tsai, and B. Schmandt (2014), 3-D crustal structure of the western United States: Application of Rayleigh-wave ellipticityextracted from noise cross-correlations, Geophys. J. Int., 198(2), 656–670, doi:10.1093/gji/ggu160.

Liu, J., J. Han, and W. S. Fyfe (2001), Cenozoic episodic volcanism and continental rifting in northeast China and possible link to Japan Seadevelopment as revealed from K–Ar geochronology, Tectonophysics, 339(3), 385–401.

Liu, Z., F. Niu, Y. J. Chen, S. Grand, H. Kawakatsu, J. Ning, et al. (2015), Receiver function images of the mantle transition zone beneath NEChina: New constraints on intraplate volcanism, deep subduction and their potential link, Earth Planet. Sci. Lett., 412, 101–111,doi:10.1016/j.epsl.2014.12.019.

Masters, G., M. Barmine, and S. Kientz (2007),Mineos User’s Manual in Computational Infrastructure for Geodynamics, Calif. Inst. of Technol, Pasadena.McKenzie, D. (1978), Some remarks on the development of sedimentary basins, Earth Planet. Sci. Lett., 40, 25–32, doi:10.1016/0012-821X(78)90071-7.Mira, A. (2001), On Metropolis-Hasting algorithms with delayed rejection, Metron, 59(3–4), 231–241.Shapiro, N. M., M. Campillo, L. Stehly, and M. H. Ritzwoller (2005), High-resolution surface-wave tomography from ambient seismic noise,

Science, 307(5715), 1615–1618, doi:10.1126/science.1108339.Shen, W., M. H. Ritzwoller, V. Schulte-Pelkum, and F.-C. Lin (2013), Joint inversion of surface wave dispersion and receiver functions:

A Bayesian Monte-Carlo approach, Geophys. J. Int., 192(2), 807–836, doi:10.1093/gji/ggs050.Smith, W. H. F., and P.Wessel (1990), Griddingwith continuous curvature splines in tension,Geophysics, 55, 293–305, doi:10.1126/science.1108339.Tang, Y., M. Obayashi, F. Niu, S. P. Grand, Y. J. Chen, H. Kawakatsu, S. Tanaka, J. Ning, and J. F. Ni (2014), Changbaishan volcanism in northeast

China linked to subduction-induced mantle upwelling, Nat. Geosci., 7(6), 470–475, doi:10.1038/ngeo2166.Tanimoto, T., and C. Alvizuri (2006), Inversion of the HZ ratio of microseisms for S-wave velocity in the crust, Geophys. J. Int., 165, 323–335.Tanimoto, T., and L. Rivera (2008), The ZH ratio method for long-period seismic data: Sensitivity kernels and observational techniques,

Geophys. J. Int., 172(1), 187–198, doi:10.1111/j.1365-246X.2007.03609.x.Tao, K., F. Niu, J. Ning, Y. J. Chen, S. Grand, H. Kawakatsu, S. Tanaka, M. Obayashi, and J. Ni (2014), Crustal structure beneath NE China imaged

by NECESSArray receiver function data, Earth Planet. Sci. Lett., 398, 48–57, doi:10.1016/j.epsl.2014.04.043..Waldhauser, F., R. Lippitsch, E. Kissling, and J. Ansorge (2002), High-resolution teleseismic tomography of upper-mantle structure using an a

priori three-dimensional crustal model, Geophys. J. Int., 150(2), 403–414, doi:10.1046/j.1365-246X.2002.01690.x.Wei, H.-H., J.-L. Liu, and Q. Meng (2010), Structural and sedimentary evolution of the southern Songliao Basin, northeast China, and

implications for hydrocarbon prospectivity, Am. Assoc. Pet. Geol. Bull., 94(4), 533–566, doi:10.1306/09080909060.Wu, F., B. Jahn, S. A. Wilde, C. Lo, T. Yui, Q. Lin, W. Ge, and D. Sun (2003a), Highly fractionated I-type granites in NE China (I): Geochronology

and petrogenesis, Lithos, 66(3), 241–273.Wu, F., B. Jahn, S. A. Wilde, C. Lo, T. Yui, Q. Lin, W. Ge, and D. Sun (2003b), Highly frac-tionated I-type granites in NE China (II): Isotopic

geochemistry and implications for crustal growth in the Phanerozoic, Lithos, 67(3–4), 191–204, doi:10.1016/S0024-4937(03)00015-X.Yang, Y., M. H. Ritzwoller, A. L. Levshin, and N. M. Shapiro (2007), Ambient noise Rayleigh wave tomography across Europe, Geophys. J. Int.,

168(1), 259–274, doi:10.1111/j.1365-246X.2006.03203.x.Yano, T., T. Tanimoto, and L. Rivera (2009), The ZH ratio method for long-period seismic data: Inversion for S-wave velocity structure, Geophys.

J. Int., 179(1), 413–424, doi:10.1111/j.1365-246X.2009.04293.x.

Journal of Geophysical Research: Solid Earth 10.1002/2015JB012459

LI ET AL. RAYLEIGH WAVE ELLIPTICITY AND S VELOCITY 17