12
Constraining Sedimentary Structure Using Frequency- Dependent P Wave Particle Motion: A Case Study of the Songliao Basin in NE China Yifei Bao 1,2 and Fenglin Niu 1,2 1 State Key Laboratory of Petroleum Resource and Prospecting, and Unconventional Gas Institute, China University of Petroleum at Beijing, Beijing, China, 2 Department of Earth Science, Rice University, Houston, TX, USA Abstract Knowledge on sedimentary structure within a basin is of great importance for exploring petroleum resources and predicting strong ground motions caused by earthquakes. Although 3-D seismic data acquisition is common nowadays in petroleum industry, it is still impossible to conduct dense 3-D seismic surveys at a whole-basin scale. Passive seismic imaging thus can play a signicant role in obtaining large-scale sedimentary structure of basins with a reasonably good resolution. Here we present a new method to estimate the sedimentary structure beneath a seismic station using frequency-dependent nonlinear particle motion of the teleseismic P wave, that is, a delay in the P wave arrival time on the horizontal component. Forward modeling suggests that the delay time and its variations over frequency are caused by interference of the direct P wave with the P-to-S conversions at the base of the sediments as well as sediment reverberations, and hence can be used to constrain the thickness and seismic velocity of the sediments beneath the station. We further develop a 2-D grid search technique to estimate the optimum sediment thickness (Z) and surface S wave velocity (β 0 ) by minimizing the residuals between the observed and calculated apparent P wave splitting times. The grid search yields reasonably robust estimates of Z and β 0 , although it has some trade-off between the two parameters in the case of thin and unconsolidated sediments. We apply the method to 30+ broadband stations inside the Songliao basin, and the estimated sedimentary structures beneath these stations agree well with previous results. 1. Introduction Sedimentary basins are formed through inlling of accommodation space created by long-term subsidence, which are the result of various tectonic processes, such as lithospheric stretching (e.g., McKenzie, 1978). Knowing sedimentary structure within a basin is also of great importance to predict strong ground motions caused by earthquakes (e.g., Fletcher & Wen, 2005; Lee, Chen, & Huang, 2008; Olsen, Archuleta, & Matarese, 1995; Taborda & Bielak, 2013). Unconsolidated sediments are formed by loose materials, ranging from clay to sand to gravel, and are usually featured by extremely low seismic velocity. Seismic waves entering the low- velocity sedimentary layer can be amplied and trapped, resulting in large ground shaking that can post devastating damage to buildings and structures. The long-lasting large-amplitude trapped waves or sedi- ment reverberations also interfere with other seismic arrivals, making it difcult to isolate signals associated with deep structures from seismic records (e.g., Langston, 2011; Owens & Crosson, 1988; Tao et al., 2014). Sedimentary basins are also known as the locations of hydrocarbon reservoirs. Active source seismic surveys have been widely conducted in basins for oil and gas exploration. While 2-D reection proles were the main- stream of early days, 3-D seismic survey really took a central role in the exploration and development of oil and gas elds over the last two decades. The spatial coverage of 2-D/3-D seismic data is, however, still very limited as compared to whole-basin scale because of the high cost of data acquisition. Thus, there is still a demand to build whole-basin-scale 3-D sedimentary models with intermediate-depth and lateral resolution in order to outline the general areas of source reservoirs for 2-D/3-D active exploration. Our goal is to develop a technique to build such 3-D models using passive seismic data collected by dense and large-scale arrays of broadband sensors. Numerous studies have been conducted to use ambient noise or microtremor to investigate the site response, also known as the transfer function, beneath a seismic station (e.g., Field & Jacob, 1993; Field et al., 1995; Lachet et al., 1996; Nakamura, 1989; Seht & Wohlenberg, 1999). These studies found peaks in the spectral ratio of the horizontal and vertical components recorded at a seismic station, referred as BAO AND NIU SEDIMENT CONSTRAINED BY P WAVE SPLITTING 1 PUBLICATION S Journal of Geophysical Research: Solid Earth RESEARCH ARTICLE 10.1002/2017JB014721 Key Points: Teleseismic P waves recorded at broadband stations above sediments exhibit strong frequency-dependent apparent splitting The apparent P wave splitting times can be used to estimate sediment thickness and surface S wave velocity Correspondence to: F. Niu, [email protected] Citation: Bao, Y., & Niu, F. (2017). Constraining sedimentary structure using frequency- dependent P wave particle motion: A case study of the Songliao Basin in NE China. Journal of Geophysical Research: Solid Earth, 122. https://doi.org/10.1002/ 2017JB014721 Received 18 JUL 2017 Accepted 16 OCT 2017 Accepted article online 19 OCT 2017 ©2017. American Geophysical Union. All Rights Reserved.

Constraining Sedimentary Structure Using Frequency

  • Upload
    others

  • View
    12

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Constraining Sedimentary Structure Using Frequency

Constraining Sedimentary Structure Using Frequency-Dependent P Wave Particle Motion: A Case Studyof the Songliao Basin in NE ChinaYifei Bao1,2 and Fenglin Niu1,2

1State Key Laboratory of Petroleum Resource and Prospecting, and Unconventional Gas Institute, China University ofPetroleum at Beijing, Beijing, China, 2Department of Earth Science, Rice University, Houston, TX, USA

Abstract Knowledge on sedimentary structure within a basin is of great importance for exploringpetroleum resources and predicting strong ground motions caused by earthquakes. Although 3-D seismicdata acquisition is common nowadays in petroleum industry, it is still impossible to conduct dense 3-Dseismic surveys at a whole-basin scale. Passive seismic imaging thus can play a significant role in obtaininglarge-scale sedimentary structure of basins with a reasonably good resolution. Here we present a newmethod to estimate the sedimentary structure beneath a seismic station using frequency-dependentnonlinear particle motion of the teleseismic Pwave, that is, a delay in the Pwave arrival time on the horizontalcomponent. Forward modeling suggests that the delay time and its variations over frequency are caused byinterference of the direct Pwave with the P-to-S conversions at the base of the sediments as well as sedimentreverberations, and hence can be used to constrain the thickness and seismic velocity of the sedimentsbeneath the station. We further develop a 2-D grid search technique to estimate the optimum sedimentthickness (Z) and surface S wave velocity (β0) by minimizing the residuals between the observed andcalculated apparent P wave splitting times. The grid search yields reasonably robust estimates of Z and β0,although it has some trade-off between the two parameters in the case of thin and unconsolidatedsediments. We apply the method to 30+ broadband stations inside the Songliao basin, and the estimatedsedimentary structures beneath these stations agree well with previous results.

1. Introduction

Sedimentary basins are formed through infilling of accommodation space created by long-term subsidence,which are the result of various tectonic processes, such as lithospheric stretching (e.g., McKenzie, 1978).Knowing sedimentary structure within a basin is also of great importance to predict strong ground motionscaused by earthquakes (e.g., Fletcher & Wen, 2005; Lee, Chen, & Huang, 2008; Olsen, Archuleta, & Matarese,1995; Taborda & Bielak, 2013). Unconsolidated sediments are formed by loose materials, ranging from clay tosand to gravel, and are usually featured by extremely low seismic velocity. Seismic waves entering the low-velocity sedimentary layer can be amplified and trapped, resulting in large ground shaking that can postdevastating damage to buildings and structures. The long-lasting large-amplitude trapped waves or sedi-ment reverberations also interfere with other seismic arrivals, making it difficult to isolate signals associatedwith deep structures from seismic records (e.g., Langston, 2011; Owens & Crosson, 1988; Tao et al., 2014).

Sedimentary basins are also known as the locations of hydrocarbon reservoirs. Active source seismic surveyshave been widely conducted in basins for oil and gas exploration. While 2-D reflection profiles were themain-stream of early days, 3-D seismic survey really took a central role in the exploration and development of oiland gas fields over the last two decades. The spatial coverage of 2-D/3-D seismic data is, however, still verylimited as compared to whole-basin scale because of the high cost of data acquisition. Thus, there is still ademand to build whole-basin-scale 3-D sedimentary models with intermediate-depth and lateral resolutionin order to outline the general areas of source reservoirs for 2-D/3-D active exploration. Our goal is to developa technique to build such 3-D models using passive seismic data collected by dense and large-scale arrays ofbroadband sensors.

Numerous studies have been conducted to use ambient noise or microtremor to investigate the siteresponse, also known as the transfer function, beneath a seismic station (e.g., Field & Jacob, 1993; Fieldet al., 1995; Lachet et al., 1996; Nakamura, 1989; Seht & Wohlenberg, 1999). These studies found peaksin the spectral ratio of the horizontal and vertical components recorded at a seismic station, referred as

BAO AND NIU SEDIMENT CONSTRAINED BY P WAVE SPLITTING 1

PUBLICATIONSJournal of Geophysical Research: Solid Earth

RESEARCH ARTICLE10.1002/2017JB014721

Key Points:• Teleseismic P waves recorded atbroadband stations above sedimentsexhibit strong frequency-dependentapparent splitting

• The apparent P wave splitting timescan be used to estimate sedimentthickness and surface S wave velocity

Correspondence to:F. Niu,[email protected]

Citation:Bao, Y., & Niu, F. (2017). Constrainingsedimentary structure using frequency-dependent P wave particle motion: Acase study of the Songliao Basin in NEChina. Journal of Geophysical Research:Solid Earth, 122. https://doi.org/10.1002/2017JB014721

Received 18 JUL 2017Accepted 16 OCT 2017Accepted article online 19 OCT 2017

©2017. American Geophysical Union.All Rights Reserved.

Page 2: Constraining Sedimentary Structure Using Frequency

to H/V spectrum, which can be explained by resonance of trapped waves inside the low-velocity sedimentarylayer. The resonant frequency is quantitatively related to the S wave velocity and thickness of the sedimentlayer (Seht & Wohlenberg, 1999), which determines the site amplification factor, a key parameter for stronggroundmotion calculation. This method works well with stations located above thin (<1 km) unconsolidatedsediments but is less efficient in resolving thick sedimentary structure.

It has been shown by previous studies (e.g., Boore & Toksöz, 1969; Chong, Ni, & Zhao, 2014) that Rayleigh-wave ellipticity or Z/H ratio is more sensitive to shallow structure than phase velocity dispersion data, whichmeans a joint inversion of phase velocity and Z/H ratio data can better constrain the near-surface shear velo-city structure. Li et al. (2016) jointly inverted the Z/H ratio and phase velocity dispersion data measured fromambient noise-based Rayleigh wave Green’s functions to obtain a 3-D S wave velocity model beneath NEChina. They found strong low-velocity anomalies at shallow depth (0–7 km) inside the Songliao basin.Since they used surface wave in the period band of 8–25 s, it is almost impossible to resolve the thin topsoillayer (<500 m).

The receiver function technique is perhaps the most widely used method in passive seismology to estimatedepth of subsurface boundaries beneath a seismic station. Due to the shallow depth of sediment base, theP-to-S conversion and sediment reverberations arrive within the P wave source time window, causing stronginterference among those phases such that it is practically impossible to employ receiver function techniqueto locate the bottom of the sediment. On the other hand, the raypaths of the conversion and reverberationwaves are nearly vertical because of the extremely low S wave velocity of the sediment. Consequently, theselater arrivals are mostly shown on the horizontal components, and the phase interference is anticipated tooccur primarily on the horizontal records.

We notice that the primary P arrival times have a discrepancy between horizontal and vertical componentsfor stations located inside sedimentary basins while they should theoretically be the same since P wave isa body wave with a linear particle motion. This discrepancy leads to a P wave delay in the receiver functiondata generated from the two components, which has been observed by previous receiver functions studies(e.g., Chen & Niu, 2016; Wang et al., 2017). The P wave delay is likely caused by interference between thedirect P and P-to-S conversion at the sediment base. Wang et al. (2017) used the Pwave delay time of receiverfunctions to study the sedimentary structure beneath the northeastern margin of the Tibetan Plateau. Wealso find that the apparent P wave (hereafter referred to as AP) splitting or nonlinear P wave particle motionvaries with frequency. The frequency-dependent AP splitting is caused by interference of the direct P wavewith P-to-S conversion at the sediment base and sediment reverberations. Thus, the AP splitting times canbe used to constrain the sedimentary structure beneath a seismic station.

In this study, we first introduce a method to measure the AP splitting time at several period bands. We furtherdevelop a grid search technique to obtain the optimum sediment thickness and S wave velocity from theobserved splitting time data. We apply our method to a large-scale broadband array temporarily deployedinside the Songliao Basin and its surrounding area, which is known as the NECESSArray (NorthEast ChinaExtended SeiSmic Array) (Tao et al., 2014) to illustrate the effectiveness of the method. The Songliao basinin northeast China is the largest nonmarine oilfield in China, which consists of six structural units: (1) thenorthern plunge, (2) the central downwarp, (3) the northeastern uplift, (4) the southeastern uplift, (5) thesouthwestern uplift, and (6) the western slope (Figure 1). Most of the petroleum reservoirs are located inthe central downwarp (Feng et al., 2010), which also appears to possess the thickest sediments inside thebasin based on our inversion.

2. The NECESSArray Data

We use teleseismic waveform data recorded by the NECESSArray, which is a temporary Program for ArraySeismic Studies of the Continental Lithosphere/ERI broadband array deployed between September 2009and August 2011 under an international collaborative project. It consists of 127 temporary broadband sta-tions with an averaged station spacing of approximately 80 km (Figure 1) that covers most part of NEChina. The Songliao basin, about 750 km long and 330–370 km wide, is located at the center of the NEChina. It is surrounded by the Great Xing’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).

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014721

BAO AND NIU SEDIMENT CONSTRAINED BY P WAVE SPLITTING 2

Page 3: Constraining Sedimentary Structure Using Frequency

We use the same teleseismic data of the receiver function studies of Tao et al. (2014) and Liu et al. (2015).More specifically, we first visually examine all the teleseismic data within the epicentral distance range of30°–90° from earthquakes with a magnitude between 5.5 and 7.5 occurring from September 2009 toAugust 2011. We then choose a total of 482 earthquakes that are well recorded by the NECESSArray stations.These earthquakes provide a good coverage in both distance and azimuth (Figure 1, inset).

3. Methodology3.1. Frequency-Dependent AP Splitting Time Measurements

For a teleseismic recording in the epicentral distance range of 30°–90°, we first rotate the two horizontal com-ponents of the seismograms to the radial and transverse components. We then band-pass filter the verticaland horizontal components with a two-pole Butterworth filter in five difference period bands: 1-10s, 2-20s,3-30s, 4-40s, and 5-50s. Figure 2 shows an example of the band-pass-filtered seismograms at 1-10s, 3-30s,and 5-50s recorded at the NE68 station, and their P wave particle motions, which exhibit large variationsamong different period bands.

We select data with a signal-to-noise ratio ≥5 and employ a cross correlation-based method to measure thetime lag of the Pwave arrival on the radial component. The cross correlation is computed in the time domainby sliding the radial component to obtain the maximum cross correlation coefficient. For station NE68, thenumber of earthquakes selected for measurement is 252 at the shortest period band (1-10s) and 142 at

116˚ 120˚ 124˚ 128˚ 132˚40˚

44˚

48˚

Gre

at X

ing’

an R

ange

Lesser Xing’an Range

F1

F2

Daqing

Sino-Korean Craton North Korea

Japan Sea

S1

-8 -4 0 4 8Bathymetry and Topography (km)

Sanjiang Basin

Yanshanorogenic belt

Jiamusi Massif

1.NP

2.CD

4.SEU

3.NEU

5.SWU

6.WS

Songliao Basin

Changbai mountain range

Zha

nggu

angc

ai R

ange

NE68

NE96

-180˚

-150˚

-120˚

-90˚

-60˚

-30˚0˚

30˚

60˚

90˚

120˚

150˚

116˚ 120˚ 124˚ 128˚ 132˚40˚

44˚

48˚

Gre

at X

ing’

an R

ange

Lesser Xing’an Range

F1

F2

Daqing

Sino-Korean Craton North Korea

Japan Sea

S1

Bathymetry and Topography (km)

Sanjiang Basin

Yanshanorogenic belt

Jiamusi Massif

1.NP

2.CD

4.SEU

3.NEU

5.SWU

6.WS

Songliao Basin

Changbai mountain range

Zha

nggu

angc

ai R

ange

NE68

NE96

-180˚

-150˚

-120˚

-90˚

-60˚

-30˚0˚

30˚

60˚

90˚

120˚

150˚

Figure 1. Map showing the 127 NECESSArray stations (solid squares), together with the Solonker suture (S1, red solid line)and two major faults (black solid lines, F1: Jiamusi-Yitong fault, F2: Dunhua-Mishan fault). The purple solid line outlines theSongliao basin, which is composed of six structural units bounded by dashed white lines: (1) NP: the northern plunge,(2) CD: the central downwarp, (3) NEU: the northeastern uplift, (4) SEU: the southeastern uplift, (5) SWU: the southwesternuplift, and (6) WS: the western slope. The black and yellow squares are stations inside the Songliao basin that showsignificant AP splitting times. The yellow solid circle in the middle denotes the Daqing city, where China’s largest oil field islocated. Inset shows the distribution of the 482 earthquakes (red stars) used in this study. The blue triangle indicates thecenter of the seismic array. Note that although the earthquakes seem to be ~20°–100° away from the array center, weonly choose stations in the epicentral distance range of 30 to 90°.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014721

BAO AND NIU SEDIMENT CONSTRAINED BY P WAVE SPLITTING 3

Page 4: Constraining Sedimentary Structure Using Frequency

the longest period band (5-50s). The AP splitting times measured from different earthquakes appear to bestable and are independent of epicentral distance (Figure 2g) and back azimuth (Figure 2h). The averagesplitting time is 0.52 ± 0.14 s at the1-10s band and decays to 0.08 ± 0.07 s at 5-50s (Figure 2i).

3.2. Forward Modeling

Tao et al. (2014) found that the crustal structure beneath NE68 is relatively simple and can be represented bya thin (~300 m) soft sedimentary cover (with a constant S wave velocity of 0.5 km/s) lying upon a moderatelythick (34.4 km) layer of crystalline bedrock (Table 1, hereafter referred to as thinSM). We use this model andemploy the Thomson-Haskell propagator matrix method (Haskell, 1962; Thomson, 1950) to generate verticaland radial seismograms of all the earthquakes. We then apply the same band-pass filters to filter the syntheticseismograms and make AP splitting time measurements with the cross-correlation-based technique. Themeasured AP splitting times are shown in Figure 2i (blue open triangles), which match the observed delay

0.0

0.4

0.8

40 60 80

120 180 240 300 3600 60

1-10s 2-20s 3-30s 4-40s 5-50s

0.0

0.4

0.8

0.0

0.4

0.8

Epicentral distance (deg.)

Back azimuth (deg.)

Period band

Split

ting

time

(s)

Split

ting

time

(s)

AP

split

ting

time

(s)BHZ

BHR

BH

Z

Time after P (s)-5 0 5 10 15

0.5

1.0

-0.5

-1.0

0.0

0.5 1.0-0.5-1.0 0.0

BHR

BH

Z

0.5

1.0

-0.5

-1.0

0.0

0.5 1.0-0.5-1.0 0.0

BHR

BH

Z

0.5

1.0

-0.5

-1.0

0.0

0.5 1.0-0.5-1.0 0.0

0.5

1.0

-0.5

0.0

0.5

1.0

-0.5

-1.0

0.0

0.5

1.0

-0.5

0.0

0.5

1.0

-0.5

-1.0

0.0

0.5

1.0

-0.5

0.0

0.5

1.0

-0.5

-1.0

0.0

Time after P (s)-5 0 5 10 15

Time after P (s)-5 0 5 10 15

BHR

BHZ

BHR

BHZ

BHR

Calculated

Observed

(a) (b)

(c) (d)

(e) (f)

(g)

(h)

(i)

NE68

1-10s

1-10s 1-10s

3-30s 3-30s

5-50s 5-50s

1-10s

1-10s

NE68 02/15/2010

Figure 2. (a) Normalized vertical- (BHZ) and radial-component (BHR) recordings of NE68 from a teleseismic earthquakeoccurring on 15 February 2010, which is filtered in the period band of 1-10s. (b) The particle motion of the P wave,which is denoted by the shaded time window in Figure 2a. (c and d) Similar to Figures 2a and 2b but in the period band of3-30s. (e and f) Similar to Figures 2a and 2b but for the period band of 5-50s. Note the strong nonlinear particle motion inthe 1-10s (Figure 2b), which changes gradually to nearly linear at 5-50s (Figure 2f). (g) Measured AP splitting times of 252teleseismic events in the period band of 1-10s are plotted against epicentral distances. The splitting times are averagedones in a 10-degree bin with error bars indicating the standard deviations. (h) The estimated AP splitting times are shown asa function of event back azimuth. The data are binned in a 20-degree window. (i) The average AP splitting time of eachperiod band is plotted as a function of period (black solid squares). The open triangles represent the computed splittingtimes from the thinSM model listed in Table 1. Note that the splitting time decays with increasing period.

Table 1Models Used in Synthetic Tests

Model H (km) α (km/s) β (km/s) ρ (g/cm3)

thinSM Sediment 0.30 2.10 0.50 1.97Crust/mantle 35.00/∞ 6.40/8.00 3.68/4.50 2.70/3.30

thickSM Sediment 4.00 equation (1) 0.68 + 0.57za equation (2)Crust/mantle 35.0/∞ 6.40/8.00 3.68/4.50 2.70/3.30

aZ is the depth in kilometer.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014721

BAO AND NIU SEDIMENT CONSTRAINED BY P WAVE SPLITTING 4

Page 5: Constraining Sedimentary Structure Using Frequency

times very well. This implies that apparent Pwave delay in the radial component is caused by the interferenceof the direct Pwave, P-to-S converted wave at the base of the sediment, and sediment reverberations. Due tothe low S wave velocity inside the sediment, the incident angle of the P-to-S conversion and reverberation isclose to zero, which means they are recorded predominately by the radial component and have almost noeffect on the vertical recordings.

The close relationship between the AP splitting times and sedimentary structure is further illustrated by twoother examples. Station NE96 is located near the city Daqing (Figure 1), which is named after the China’s lar-gest oil field, Daqing (also known as Taching) Oil Field inside the central downwarp of the Songliao basin. ThePwaves recorded by the radial component of the station, which is covered by very thick sediments (reaching

(a) (b)

(c)

BHZ

BHR

BH

Z

Time after P (s)-5 0 5 10 15 0.5 1.0-0.5-1.0 0.0

0.5

1.0

-0.5

0.0

0.5

1.0

-0.5

-1.0

0.0

BHR

1-10s 1-10s

0.5

1.0

-0.5

-1.0

0.0

BHZ

BHR

BH

Z

Time after P (s)-5 0 5 10 15 0.5 1.0-0.5-1.0 0.0

0.5

1.0

-0.5

0.0

0.5

1.0

-0.5

-1.0

0.0

BHR

5-50s 5-50s

0.5

1.0

-0.5

-1.0

0.0

1-10s 2-20s 3-30s 4-40s 5-50s

0.0

1.5

2.0

Period band Sp

littin

g tim

e (s

)1-10s

-0.5

1.0

0.5

(d)

(e)

NEA3

NE96 02/15/2010

NE96

Observed AP spliting times

Figure 3. (a) Normalized vertical- (BHZ) and radial-component (BHR) recordings of NE96 from a teleseismic earthquakeoccurring on 15 February 2010, which is filtered in the period band of 1-10s. (b) The particle motion of the P wave, whichis denoted by the shaded time window in Figure 3a. (c and d) Similar to Figures 3a and 3b but in the period band of 5-50 s.(e) The average AP splitting times measured at NE96 are plotted as a function of period (red solid squares), whichremain flat across all the period bands, clearly different from those observed at NE68 (Figure 2i). For comparison, the APsplitting times measured at NEA3, a station deployed on bedrock, are also shown (open blue circles). The splitting timesmeasured at NEA3 are all close to zero.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014721

BAO AND NIU SEDIMENT CONSTRAINED BY P WAVE SPLITTING 5

Page 6: Constraining Sedimentary Structure Using Frequency

to several kilometers), show very large delay as compared to those on the vertical component (Figures 3aand 3c), resulting in distinct nonlinear particle motions (Figures 3b and 3d). The measured AP splittingtimes are above 1 s across all the five period bands (Figure 3e), suggesting that the amplitudes of the APsplitting time and its variations across different period bands are very sensitive to sedimentary structurebeneath the seismic station. As a comparison, station NEA3 is located on bedrock, and we find nosignificant AP splitting at all the period bands (Figure 3e).

3.3. Grid Search of Sediment Thickness and Velocity

To constrain sediment thickness and velocity from the observed AP splitting times, we develop a grid-search-based technique to search for the optimum sediment thickness and Swave velocity that match the observed

0.3

0.4

0.5

0.6

0.7

0.1 0.2 0.3 0.4 0.50.3

0.4

0.5

0.6

0.7

0.1 0.2 0.3 0.4 0.5

0.3

0.4

0.5

0.6

0.7

0.8

3.2 3.6 4.0 4.4 4.80.3

0.4

0.5

0.6

0.7

0.8

3.2 3.6 4.0 4.4 4.8

0 80 85 90 95 100Normalized variance reduction (%)

Sediment depth (km) Sediment depth (km)

Surf

ace

S-ve

loci

ty (

km/s

)Su

rfac

e S-

velo

city

(km

/s)

Sediment depth (km) Sediment depth (km)

NE96

NE68

ThickSM

ThinSM

(a)

(b)

(c)

(d)

Figure 4. (a) The 2-D grid search result computed from the synthetic model thinSM with thin and unconsolidated sedi-ment, designed to mimic the sedimentary structure beneath NE68. The normalized variance reduction is indicated bycolor contour map in which “hotter” color clusters represent greater variance reduction. The two thin white lines indicatethe optimum sediment thickness and surface S wave velocity at which the variance reduction reaches the maximum.(b) Same as Figure 4a except for the synthetic model thickSM with a thick sedimentary layer, which is observed beneathNE96. Note the large uncertainty in the thickness estimate. (c) Similar to Figure 4a but the grid search is computed from thereal AP splitting times measured at NE68. Note the multiple peaks on the color contour map, suggesting that there is atrade-off between sediment thickness and surface S wave velocity. (d) Similar to Figure 4a but the grid search is computedfrom the real AP splitting times measured at NE96. Note the large uncertainty in the measured sediment thickness, whichimplies that the AP splitting is mainly caused by shallow conversions and multiples.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014721

BAO AND NIU SEDIMENT CONSTRAINED BY P WAVE SPLITTING 6

Page 7: Constraining Sedimentary Structure Using Frequency

AP splitting times. We scale the P wave velocity (α) from S wave velocity(β) using a linear relationship complied by Castagna, Batzle, andEastwood (1985) from clastic silicate rocks:

α ¼ 1:16β þ 1:36 (1)

We compute density, ρ, from α using the scaling equation obtained byBrocher (2005):

ρ g=cm3� � ¼ 1:6612α� 0:4721α2 þ 0:0671α3 � 0:0043α4

þ 0:000106α5 (2)

We further assume that shear-wave velocity increases linearlywith depth:

β zð Þ ¼ β0 þ kz; (3)

where β0 is the S wave velocity at the surface and k is the velocity gradi-ent with respect to the depth (δβ/δz), which is assumed to be constanthere. When k is zero, velocity inside the sediment is constant. We para-meterize the sediments with a stack of constant velocity layers with athickness between 0.2 and 0.4 km. Beneath the sedimentary layer, thecrystalline crust is assumed to have a constant P and S wave velocityand density, which is 6.4 km/s, 3.46 km/s, and 2.7 g/cm3, respectively.It is underlaid by a half-space mantle with a constant α = 8.0 km/s,β = 4.5 km/s, and ρ = 3.3 g/cm3.

We vary the sediment thickness from Zmin to Zmax and β0 from β0min toβ0max with an increment of Δz = 0.02 km and Δβ = 0.01 km/s, respec-tively. The k is chosen with a try and error approach. For each searchedmodel, we use the Thomson-Haskell propagator matrix method (Haskell,1962; Thomson, 1950) to generate vertical and radial seismograms of allthe earthquakes. The incident angle is calculated using the iasp91(Kennett & Engdahl, 1991) based on the averaged epicentral distanceand focal depth. The synthetic seismograms are filtered with the same

band-pass filters. We then measure the AP splitting times from the filtered synthetic seismograms of allthe earthquakes and further compute their average. The objective function is taken as the weighted averageresidual between the observed and calculated AP splitting times:

ΔT ¼ 1N

XNi¼1

Toi � Tcið Þ2σ2oi

" #12

(4)

Here N is total number of period bands, which is 5. Toi and Tci are the observed and calculated AP splittingtime of the ith period band, and σoi is the uncertainty of Toi. The thickness and velocity ranges are chosenwhen the average time residual, ΔT, reaches minimum.

To illustrate the effectiveness of the grid search approach, we use the AP splitting times measured from thesynthetic seismograms generated from the thinSM model (shown in Figure 2i) as data to invert Z and β0. Wetake (Zmin, Zmax, and Δz) and (β0min, β0max, and Δβ) to be (0.1, 0.5, and 0.02) in kilometer and (0.3, 0.7, and 0.01)in km/s, respectively. The δβ/δz is taken to be zero. The result is shown in Figure 4a. The optimum (Z, β) is(0.3 km, 0.5 km/s), which is exactly the same as the input value. Since the P wave velocity and density of theinput model (thinSM) do not follow the scaling equations (1) and (2), therefore, their values (α = 1.94 km/s,

0.0

0.4

0.8

1.2

1.6

2-20s 3-30s 4-40s 5-50sPeriod band

1-10s

Split

ting

time

(s)

0.0

0.4

0.8

1.2

1.6

Split

ting

time

(s)

Observed AP splitting times

2-20s 3-30s 4-40s 5-50sPeriod band

1-10s

(a)

(b)

Figure 5. AP splitting times estimated from the 30 stations within theSongliao basin (black and yellow solid squares in Figure 1). (a) The “NE68”-type and “NE96”-type splitting time curves are shown in black squaresand red circles, respectively. (b) Stations with a splitting time distributedbetween the two end-members shown in Figure 5a.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014721

BAO AND NIU SEDIMENT CONSTRAINED BY P WAVE SPLITTING 7

Page 8: Constraining Sedimentary Structure Using Frequency

ρ = 1.88 g/cm3) in the final model obtained from the grid search are slightly different from the input model(α = 2.10 km/s, ρ = 1.97 g/cm3).

To simulate the large AP splitting times observed at station NE96, we create a synthetic model (hereafterreferred to as thickSM) with a sediment thickness of 4.0 km, and β0 = 0.68 km/s and δβ/δz = 0.57 km/s/km.The P wave velocity and density are computed from the scaling equations (1) and (2). We use the syntheticAP splitting times measured from the Haskell synthetics of the thickSM model as the data and conduct thegrid search. The sediment thickness range is set to be 3–5 km, and surface S wave velocity is searchedbetween 0.3 and 0.8 km/s. We employ k = 0.57 km/s/km in the grid search. The optimum thickness and sur-face S wave velocity are 4 km and 0.68 km/s, respectively, exactly the same as the input values.

4. Results4.1. Observed AP Splitting Times

We obtain high-quality AP splitting times at a total of 30 stations inside the Songliao basin (black and yel-low solid squares in Figure 1). The measurements are shown in Figure 5 and are also listed in Table 2. Theobserved AP splitting times at the 30 stations can be divided roughly into three groups: (1) those havesmall to moderate splitting times at short-period bands, which decrease sharply to zero (black symbols inFigure 5a); we refer it to as the NE68 group hereafter; (2) those have relatively large splitting times acrossall the five period bands (red symbols in Figure 5a, NE96 group); and (3) those between (Figure 5b,middle group).

Table 2Observed AP Splitting Times and 2-D Grid Search Results

Station

Lon. Lat. Elev. AP splitting time (s) δβ/δz Z β VR

(deg) (deg) (km) 1_10s 2_20s 3_30s 4_40s 5_50s a (km) (km/s) (%)

NE11 124.1 42.9 0.15 0.442 ± 0.105 0.422 ± 0.097 0.359 ± 0.092 0.291 ± 0.086 0.287 ± 0.107 1.50 2.10 0.60 74.57NE38 122.4 42.8 0.26 0.307 ± 0.104 0.176 ± 0.080 0.115 ± 0.085 0.100 ± 0.098 0.148 ± 0.127 1.50 1.60 0.98 61.83NE39 123.3 43.0 0.14 0.269 ± 0.114 0.099 ± 0.064 0.036 ± 0.042 0.043 ± 0.052 0.080 ± 0.078 0.10 0.28 0.78 69.58NE45 120.6 43.4 0.31 0.423 ± 0.094 0.297 ± 0.080 0.230 ± 0.095 0.189 ± 0.114 0.221 ± 0.148 1.50 2.00 0.72 70.55NE46 121.5 43.4 0.23 1.074 ± 0.132 0.972 ± 0.098 0.878 ± 0.097 0.791 ± 0.089 0.734 ± 0.074 0.60 5.22 0.78 92.72NE47 124.3 43.5 0.18 0.501 ± 0.144 0.546 ± 0.155 0.435 ± 0.144 0.338 ± 0.134 0.314 ± 0.119 1.50 2.00 0.50 80.38NE55 121.6 44.2 0.21 0.672 ± 0.128 0.541 ± 0.138 0.391 ± 0.140 0.274 ± 0.123 0.273 ± 0.143 1.50 2.00 0.50 83.53NE56 122.4 44.1 0.17 0.626 ± 0.151 0.509 ± 0.150 0.351 ± 0.127 0.233 ± 0.113 0.191 ± 0.106 1.50 1.80 0.54 86.85NE57 123.3 44.1 0.15 1.172 ± 0.107 1.061 ± 0.104 0.946 ± 0.112 0.835 ± 0.141 0.705 ± 0.124 0.60 5.22 0.74 89.88NE58 125.3 44.4 0.20 0.473 ± 0.113 0.490 ± 0.102 0.464 ± 0.136 0.447 ± 0.170 0.454 ± 0.183 0.60 3.00 1.00 71.63NE68 122.8 44.9 0.15 0.522 ± 0.138 0.354 ± 0.126 0.182 ± 0.089 0.096 ± 0.072 0.074 ± 0.071 0.00 0.30 0.50 95.90NE69 123.7 44.8 0.16 1.343 ± 0.086 1.276 ± 0.120 1.227 ± 0.128 1.165 ± 0.122 1.073 ± 0.112 0.60 5.22 0.60 97.52NE6A 124.5 44.8 0.18 0.733 ± 0.080 0.705 ± 0.083 0.674 ± 0.096 0.601 ± 0.097 0.559 ± 0.099 0.60 4.82 0.96 92.27NE6C 127.6 44.9 0.21 0.451 ± 0.210 0.389 ± 0.153 0.297 ± 0.107 0.257 ± 0.100 0.275 ± 0.118 1.50 2.00 0.62 73.17NE78 123.2 45.5 0.14 0.751 ± 0.126 0.627 ± 0.140 0.450 ± 0.139 0.295 ± 0.121 0.234 ± 0.108 1.50 1.12 0.40 87.81NE79 124.1 45.5 0.13 1.166 ± 0.171 1.145 ± 0.104 1.149 ± 0.082 1.124 ± 0.078 1.114 ± 0.091 0.40 3.84 0.84 92.76NE7A 126.0 45.5 0.15 0.530 ± 0.153 0.580 ± 0.095 0.584 ± 0.099 0.595 ± 0.109 0.638 ± 0.137 0.60 3.34 1.00 82.42NE7B 126.9 45.5 0.18 0.397 ± 0.157 0.296 ± 0.174 0.234 ± 0.190 0.319 ± 0.312 0.384 ± 0.305 1.50 2.00 0.68 50.34NE88 123.3 46.2 0.14 0.335 ± 0.126 0.206 ± 0.094 0.125 ± 0.070 0.090 ± 0.071 0.104 ± 0.097 1.50 1.18 0.90 74.67NE89 124.2 46.2 0.14 1.233 ± 0.207 1.090 ± 0.110 1.029 ± 0.079 1.032 ± 0.079 1.087 ± 0.102 0.60 5.36 0.66 94.80NE8A 125.1 46.2 0.14 1.202 ± 0.170 1.028 ± 0.099 1.056 ± 0.092 1.090 ± 0.094 1.078 ± 0.100 0.60 3.38 0.64 96.23NE8B 126.1 46.2 0.17 0.787 ± 0.121 0.748 ± 0.154 0.641 ± 0.170 0.530 ± 0.158 0.475 ± 0.136 0.60 4.82 0.96 86.85NE8C 126.9 46.2 0.16 0.590 ± 0.173 0.611 ± 0.159 0.515 ± 0.155 0.420 ± 0.122 0.407 ± 0.143 1.50 2.00 0.44 79.34NE94 123.3 46.9 0.15 0.335 ± 0.083 0.201 ± 0.073 0.162 ± 0.090 0.118 ± 0.094 0.121 ± 0.111 1.50 1.64 0.90 71.72NE95 124.2 46.9 0.14 1.101 ± 0.117 1.100 ± 0.122 1.089 ± 0.117 0.978 ± 0.122 0.871 ± 0.128 0.60 4.82 0.68 94.00NE96 125.1 46.9 0.15 1.152 ± 0.101 1.128 ± 0.088 1.080 ± 0.083 1.052 ± 0.083 1.075 ± 0.096 0.57 4.02 0.64 95.02NE98 126.6 47.3 0.19 0.422 ± 0.113 0.269 ± 0.119 0.174 ± 0.138 0.160 ± 0.169 0.200 ± 0.195 1.50 2.00 0.78 67.34NEA5 124.0 47.6 0.16 0.962 ± 0.148 0.904 ± 0.179 0.727 ± 0.209 0.496 ± 0.200 0.351 ± 0.157 0.10 0.36 0.36 95.41NEA6 125.0 47.6 0.16 1.022 ± 0.238 1.046 ± 0.239 1.055 ± 0.205 0.929 ± 0.173 0.838 ± 0.179 0.60 4.44 0.72 92.93NEA7 125.8 47.6 0.23 0.706 ± 0.258 0.784 ± 0.290 0.672 ± 0.298 0.525 ± 0.273 0.393 ± 0.248 1.50 3.20 0.34 90.08

akm/s/km.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014721

BAO AND NIU SEDIMENT CONSTRAINED BY P WAVE SPLITTING 8

Page 9: Constraining Sedimentary Structure Using Frequency

Most stations in the NE96 group are located in the central downwarpsection of the Songliao basin (Figure 6), while those in the NE68 andthe middle groups belong to the other geological units (Figure 6). Weuse a linear regression to measure the slope of the splitting time-frequency dependence and find that most of the stations have a nega-tive slope (shown in red “dash“ signs in Figure 6). Some stations in theNE96 group, on the other hand, have an almost flat distribution of split-ting times across all the period bands (shown in black cross in Figure 6).

4.2. Grid Search Results

For each station, we employ three thickness/velocity ranges andvelocity-depth slopes with a thin, moderately thick, and thick sedimen-tary layer, respectively, in searching the optimum sedimentary structurebeneath each station. The obtained thickness and surface S wave velo-city are listed in Table 2. The variance reduction varies from 50.34% to97.52% with an average value of 83.07%. The grid search result ofNE68 is shown in Figure 4c, and the measured sediment thickness andsurface S wave velocity are 0.3 km and 0.5 km/s, respectively. Thesevalues agree well with those derived from the wavefield-downward-continuation method (Tao et al., 2014). Figure 4d shows the grid searchresult of YP.N96, where the optimum sediment thickness is 4.24 km, andthe surface S wave velocity is 0.68 km/s. Both are consistent with theresults from the surface wave study (Li et al., 2016).

We compare our results with those derived from the joint inversion ofRayleigh wave phase velocity dispersion and Rayleigh wave ellipticity(Z/H amplitude ratio) data (Li et al., 2016). To make quantitative compar-ison, we take the depth where S wave velocity reaches to 2.5 km/s assedimentary thickness (Z2.5), instead of using those derived from gridsearch. In our case, we use equation (3) to compute Z2.5. If Z2.5 is largerthan the grid-search-based depth, which means that S wave velocityinside the entire sedimentary layer is less than 2.5 km/s, then we takethe grid-search-based sediment thickness. The comparison is shown inFigure 7a. The maximum depth resolution of the joint inversion (Li et al.,2016) is 1 km, which could deteriorate the correlation. In general, thetwo measurements show a positive correlation (Figure 7a), while ourmeasurements are systematically smaller than those from the jointinversion (Figure 7a). This is likely related to the fact that we employan oversimplified crustal and mantle model in the grid search. But thedifferences of the two measurements are generally less than 0.5 km.

Five stations are located very close (<30 km) to the CC0 profile shown in the Figure 11 of Li et al. (2016).We plot the Z2.5 measured at these four stations in the depth section CC0, which suggests that the two typesof measurements of sediment thickness indeed agree well with each other.

5. Discussion

As mentioned in section 1, the apparent P wave delay in the radial component and its variation with domi-nant frequency is caused by the interference of the direct P wave, P-to-S converted wave at the base ofthe sediment, and sediment reverberations. This is because the raypaths of the conversion and reverberationwaves are nearly vertical; therefore, they are mostly recorded by the horizontal components. The relative arri-val times and amplitudes of these later phases are uniquely determined by the thickness and velocity of thesediments beneath the recording stations. If a seismic station is placed above a thin and unconsolidated sedi-ment layer, then the sediment thickness and S wave velocity can be well constrained by the frequency-dependent AP splitting times, as demonstrated by the synthetic test shown in Figure 4a. In principal, the Pwave velocity of the sediment is also expected to affect the relative arrival times and amplitudes of the

AP splitting times at 1-10s

AP splitting times at 5-50s

1.2s

1.0s

0.8s

0.6s0.4s

APST

1.2s

1.0s

0.8s

0.6s0.4s

APST

(a)

(b)

C

C’

Figure 6. Map showing the AP splitting times estimated from the 30 stationswithin the Songliao basin, with (a) being measured at 1-10s and (b) at 5-50s.The size of the circles is proportional to the measured splitting times. Thered “dash” signs indicate stations with a splitting time that decreases withincreasing period, while the “cross” signs represent stations with a flatdistribution of splitting time. (a) The open black circles, solid circles with red,and blue outlines represent the NE68, NE96, and the middle groups respec-tively. (b) The yellow line labeled as CC0 shows the location of the depthsection shown in Figure 7b.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014721

BAO AND NIU SEDIMENT CONSTRAINED BY P WAVE SPLITTING 9

Page 10: Constraining Sedimentary Structure Using Frequency

converted wave and multiples. We find that a ~10% deviation from thetrue P wave model still gives accurate estimates of the sediment thick-ness and S wave velocity, and the Castagna et al.’s (1985) scaling is agood representation of the P wave velocity of the sediment. For the realdata, we found a moderate trade-off between sediment thickness and Swave velocity due to the uncertainties in the AP splitting time measure-ments. As shown in Figure 4c, multiple peaks can be found in the 2-Dgrid search domain (Z, β0). In general, these peaks are located with0.1 km of the true depth, which can be considered as the depth resolu-tion of this method and is endurable for a whole-basin scale model.

For stations located above thick sediments, we find that the AP splittingtimes are less sensitive to sediment thickness. As shown in Figure 4b, theuncertainty in the thickness estimate is around 0.1 km even with noise-free synthetic data. This is likely related to the fact that the AP splitting ismainly caused by interference from the shallow conversions and rever-berations. We also employ a slightly different α-β scaling rule for thedeep sediments where α is above 4 km/s (Brocher, 2005):

α ¼ 0:9409þ 2:0947β � 0:8206β2 þ 0:2683β3 � 0:0251β4 (5)

and find that the results remain more or less the same. When the sedi-mentary layer is thick, seismic velocity is expected to increase withdepth due to compaction and lithification. We explore a variety ofk = δβ/δz and find that the optimum k lies in the range of 0.5–0.7 km/s/km, which is consistent with field observations (e.g., Limpornpipatet al., 2012).

As mentioned in section 3, we have employed a one-layer crystallinecrust model in the grid search, which is an oversimplification and mayaffect the grid search results. To investigate how the employedcrystalline crustal and mantle model affect the grid search results, wecreate a crustal model by replacing the top 4.8 km of the S wave modelof NE96 inverted from Rayleigh wave data by Li et al. (2016) with ahypothetic sedimentary layer that has an S wave velocity increasinglinearly from β0 = 0.7 km/s and δβ/δz = 0.57 km/s/km (solid red line inFigure 8a). There are more than 10 layers below, which are categor-ized as the crystalline crust. We then use a single-layer crystalline crust(SLCC) model and the above multilayered crystalline crust (MLCC)

model to search separately for the optimum sediment thickness and surface S wave velocity. The resultsare shown in Figures 8b and 8c, respectively. The SLCC-based search yields a sediment thickness of~3.6 km, almost 1.2 km shallower than the input depth. The MLCC-based search, on the other hand, pro-duces the correct surface S wave velocity (0.7 km/s) and sediment thickness (4.8 km) with a very largeuncertainty (~ ± 0.5 km). The S wave velocity model derived from the SLCC search is also shown inFigure 8a (blue dotted line) for comparison. If we define the sediment thickness as the depth where Swave velocity reaches to 2.5 km/s, Z2.5, then the two searches yield almost the same estimate of sedimentthickness, which is approximately 3.2 km. In this regard, the crystalline crustal model used in the gridsearch has no significant effect on the results.

It is all well known that receiver functions from stations located above sediment tend to be dominated bylarge oscillations instead of isolated pulses because of the strong near-surface multiples (e.g., Owens &Crosson, 1988; Tao et al., 2014). The poor data quality is also partly related to the nature of deconvolutionused in generating receiver functions, which is a numerically unstable procedure. Bodin, Yuan, andRomanowicz (2014) proposed a Bayesian inversion with a cross-convolution misfit function to avoid theunstable deconvolution process. We have used the AP splitting time as misfit function since measurement

0

1

2

3

4

5

2

4

6

8

1043 44 45 46 47

1.6 2.0 2.4 2.8 3.2 3.6

0 1 2 3 4 5Sediment thickness from AP splitting (km)

Sedi

men

t thi

ckne

ss f

rom

join

t inv

ersi

on (

km)

Dep

th (

km)

S-wave velocity (km/s)

Latitude

C, SW C’, NE

(a)

(b)

3.2

2.8

2.4

2.02.0

1.6 1.6

2.42.8

3.2

Figure 7. (a) A comparison of sediment thickness measured from the APsplitting times and surface wave data. Note that the base of the sedimen-tary layer is defined as the surface where the S wave velocity reaches to2.5 km/s. (b) Locations of the sediment base beneath five stations are plottedon the S wave velocity map along the CC0 line across the Songliao basin. TheS wave velocity is from Li et al. (2016), which was inverted from Rayleighwave phase velocity and Z/H ratio data.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014721

BAO AND NIU SEDIMENT CONSTRAINED BY P WAVE SPLITTING 10

Page 11: Constraining Sedimentary Structure Using Frequency

of apparent splitting is relatively simple, and the grid search can be conducted very easily with a reasonabledepth resolution. However, it should be noted that the cross convolution is likely a better misfit function toconduct the grid search, which is anticipated to have better depth resolution and therefore is worth exploringfurther in future studies.

6. Conclusions

In this study, we investigate the particle motion of the teleseismic P waves recorded by the NECESSArray inNE China. We find that the teleseismic P waves recorded by stations located above sediments exhibit afrequency-dependent nonlinear particle motion, which is caused by interference from the P-to-S convertedwave at the base of the sediment and reverberations within the sediments. The observed apparent P wavedelay times on the radial component and their variations across different period bands are closely relatedto the sedimentary structure beneath the station. Sediment thickness and surface Swave velocity can be con-strained through a 2-D grid search that minimizes the apparent splitting time residuals. When sediment isthick, there is a large uncertainty in the estimate sediment thickness. The AP splitting time data can be com-bined with surface wave and receiver function data to better constrain sedimentary structure beneath broad-band seismic stations.

ReferencesBodin, T., Yuan, H., & Romanowicz, B. (2014). Inversion of receiver functions without deconvolution—Application to the Indian craton.

Geophysical Journal International, 196(2), 1025–1033. https://doi.org/10.1093/gji/ggt431Boore, D. M., & Toksöz, M. N. (1969). Rayleigh wave particle motion and crustal structure. Bulletin of the Seismological Society of America, 59(1),

331–346.Brocher, T. M. (2005). Empirical relations between elastic wavespeeds and density in the Earth’s crust. Bulletin of the Seismological Society of

America, 95(6), 2081–2092. https://doi.org/10.1785/0120050077Castagna, J. P., Batzle, M. L., & Eastwood, R. L. (1985). Relationships between compressional-wave and shear-wave velocities in clastic silicate

rocks. Geophysics, 50(4), 571–581. https://doi.org/10.1190/1.1441933Chen, Y., & Niu, F. (2016). Joint inversion of receiver functions and surface waves with enhanced preconditioning on densely distributed

CNDSN stations: Crustal and upper mantle structure beneath China. Journal of Geophysical Research: Solid Earth, 121(2), 743–766. https://doi.org/10.1002/2015JB012450

Chong, J., Ni, S., & Zhao, L. (2014). Joint inversion of crustal structure with the Rayleigh wave phase velocity dispersion and the ZH ratio.Pure and Applied Geophysics, 172(10), 2585–2600. https://doi.org/10.1007/s00024-014-0902-z

Feng, Z., Jia, C., Xie, X., Zhang, S., Feng, Z., & Cross, T. A. (2010). Tectonostratigraphic units and stratigraphic sequences of the nonmarineSongliao basin, northeast China. Basin Research, 22(1), 79–95. https://doi.org/10.1111/j.1365-2117.2009.00445.x

Field, E. H., & Jacob, K. (1993). The theoretical response of sedimentary layers to ambient seismic noise. Geophysical Research Letters, 20(24),2925–2928. https://doi.org/10.1029/93GL03054

Field, E. H., Clement, A. C., Jacob, K. H., Aharonian, V., Hough, S. E., Friberg, P. A., … Abramian, H. A. (1995). Earthquake site-responsestudy in Giumri (formerly Leninakan), Armenia, using ambient noise observations. Bulletin of the Seismological Society of America, 85,349–353.

0.5

0.6

0.7

0.8

3.6 4.0 4.4 4.8 5.20.5

0.6

0.7

0.8

4.0 4.4 4.8 5.2

Surf

ace

S-ve

loci

ty (

km/s

)

Sediment depth (km) Sediment depth (km)

MLCCSLCC

(b) (c)0

2

4

6

8

101.0 2.0 3.0 4.0

Dep

th (

km)

S-wave velocity

Z2.5ZSLCC

ZMLCC

(a)

Figure 8. Effect of the employed crystalline crust model on the 2-D grid search results. (a) The red line shows the input synthetic model, which has a 4.8 km thicksediment layer underlain by a crystalline crust taken from the NE96 velocity model of the Li et al. (2016). The blue line represents the sedimentary modelderived from the 2-D grid search using a reference model that has a single layer crystalline crust. (b) The 2-D grid search result computed with a reference modelconsisting of a sedimentary cover and a single-layer crystalline crust (SLCC). (c) Same as Figure 8b except for the reference model, which is composed of asedimentary cover and a multilayered crystalline crust (MLCC) similar to the input model.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014721

BAO AND NIU SEDIMENT CONSTRAINED BY P WAVE SPLITTING 11

AcknowledgmentsWe thank all the people involved in theNECESSArray project for providing thewaveform data, Kai Tao and Guoliang Liat the China University of Petroleum atBeijing for providing the NE68 velocitymodel and the NE96 velocity model. Wealso thank the Associate Editor and twoanonymous reviewers for their con-structive comments and suggestions,which significantly improved the qualityof this paper. All the data used in thisstudy are in public domain, which canbe downloaded from the IRIS DataManagement Center (http://ds.iris.edu/ds/nodes/dmc/). This study is supportedby NSF EAR-1547228 and NSFc41630209.

Page 12: Constraining Sedimentary Structure Using Frequency

Fletcher, J. B., & Wen, K.-L. (2005). Strong ground motion in the Taipei Basin from the 1999 Chi-Chi, Taiwan, earthquake. Bulletin of theSeismological Society of America, 95(4), 1428–1446. https://doi.org/10.1785/0120040022

Haskell, N. A. (1962). Crustal reflection of plane P and SV waves. Journal of Geophysical Research, 67(12), 4751–4768. https://doi.org/10.1029/JZ067i012p04751

Kennett, B., & Engdahl, E. R. (1991). Traveltimes for global earthquake location and phase identification. Geophysical Journal International,105(2), 429–465. https://doi.org/10.1111/j.1365-246X.1991.tb06724.x

Lachet, C., Hatzfel, P., Bard, P.-Y., Theodulidis, N., Papaioannou, C., & Savvaidis, A. (1996). Site effects and microzonation in the City ofThessaloniki (Greece) comparison of different approaches. Bulletin of the Seismological Society of America, 86, 1692–1703.

Langston, C. A. (2011). Wave-field continuation and decomposition for passive seismic imaging under deep unconsolidated sediments.Bulletin of the Seismological Society of America, 101(5), 2176–2190. https://doi.org/10.1785/0120100299

Lee, S.-J., Chen, H.-W., & Huang, B.-S. (2008). Simulations of strong ground motion and 3D amplification effect in the Taipei Basin by using acomposite grid finite-difference method. Bulletin of the Seismological Society of America, 98(3), 1229–1242. https://doi.org/10.1785/0120060098

Li, G., Chen, H., Niu, F., Guo, Z., Yang, Y., & Xie, J. (2016). Measurement of Rayleigh wave ellipticity and its application to the joint inversion ofhigh-resolution S wave velocity structure beneath northeast China. Journal of Geophysical Research, 121, 864–880. https://doi.org/10.1002/2015JB012459

Limpornpipat, O., Laird, A., Morley, C., Tingay, M., Kaewla, C., & Macintyre, H. (2012). Overpressures in the Northern Malay Basin: Part 2—Implications for pore pressure prediction. Bangkok, Thailand: International Petroleum Technology Conference.

Liu, Z., Niu, F., Chen, J. Y., Grand, S., Kawakatsu, H., Ning, J.,… Ni, J. (2015). Receiver function images of the mantle transition zone beneath NEChina: New constraints on intraplate volcanism and deep subduction. Earth and Planetary Science Letters, 412, 101–111. https://doi.org/10.1016/j.epsl.2014.12.019

McKenzie, D. (1978). Some remarks onthe development of sedimentary basins. Earth and Planetary Science Letters, 40(1), 25–32. https://doi.org/10.1016/0012-821X(78)90071-7

Nakamura, Y. (1989). A method for dynamic characteristic estimation of subsurface using microtremor on grand surface. Quarterly ReportRTRI, Japan, 30, 25–33.

Olsen, K. B., Archuleta, R. J., & Matarese, J. R. (1995). Three-dimensional simulation of a magnitude 7.75 earthquake on the San Andreas faultin southern California. Science, 270(5242), 1628–1632. https://doi.org/10.1126/science.270.5242.1628

Owens, T. J., & Crosson, R. S. (1988). Shallow structure effects on broadband teleseismic P waveforms. Bulletin of Seismological Society ofAmerica, 78, 96–108.

Seht, M. I., & Wohlenberg, J. (1999). Microtremor measurements used to map thickness of soft sediments. Bulletin of the Seismological Societyof America, 89, 250–259.

Taborda, R., & Bielak, J. (2013). Ground-motion simulation and validation of the 2008 Chino Hills, California, earthquake. Bulletin of theSeismological Society of America, 103(1), 131–156. https://doi.org/10.1785/0120110325

Tao, K., Niu, F., Ning, J., Chen, Y. J., Grand, S., Kawakatsu, H., … Ni, J. (2014). Crustal structure beneath NE China imaged by NECESSArrayreceiver function data. Earth and Planetary Science Letters, 398, 48–57. https://doi.org/10.1016/j.epsl.2014.04.043

Thomson, W. T. (1950). Transmission of elastic waves through a stratified solid medium. Journal of Applied Physics, 21(2), 89–93. https://doi.org/10.1063/1.1699629

Wang, W., Wu, J., Fang, L., Lai, G., & Cai, Y. (2017). Sedimentary and crustal thicknesses and Poisson’s ratios for the NE Tibetan Plateau and itsadjacent regions based on dense seismic arrays. Earth and Planetary Science Letters, 462, 76–85. https://doi.org/10.1016/j.epsl.2016.12.040

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014721

BAO AND NIU SEDIMENT CONSTRAINED BY P WAVE SPLITTING 12