SUMMARY

The Gonghe Basin in the northeast Tibetan Plateau presents significant potential for hot dry rock (HDR) geothermal resources. A 1990 Mw 6.4 earthquake in the basin furthers the need for an improved understanding of its sedimentary structure. In this study, we utilize data from a dense seismic array of 88 short-period seismometers deployed at an interstation spacing of approximately 3 km to scrutinize the sedimentary structure of the Gonghe Basin. By analysing teleseismic P waveforms, we identify P-to-S converted waves (Ps wave) originating from the sedimentary basement. We then determine the delay time between the Ps waves and the direct P waves (P wave) through waveform cross-correlation. By integrating this delay time with empirical velocity structure models, HDR borehole data and results from teleseismic receiver function analysis, we derive a sediment thickness model of the Gonghe Basin for the Qabqa geothermal area. Our findings reveal a gradual increase in sediment thickness from around 500 m in the east to approximately 3000 m in the west, which is consistent with other geophysical surveys and borehole data. The thick sediments in the basin could potentially serve as an excellent thermal storage cover for HDR. The strong ground motion simulation using our sediment thickness model shows that thick sediments can amplify seismic waves, increasing the risk of seismic hazards. Moreover, our study indicates that the clear Ps waves can be effectively extracted to construct a dependable sediment thickness model using teleseismic P waves recorded by a short-period dense seismic array.

1 INTRODUCTION

The Gonghe Basin, located in the northeast Tibetan Plateau and situated at the junction of the Qilian, Qaidam and Songpan-Ganzi massifs, is a rhombic graben basin that has been evolving since the Middle Cenozoic era (Fig. 1). During the Middle and Late Mesozoic periods, the basin margin experienced rupture, resulting in gradual uplift of the surrounding mountain ranges, while the central region of the basin continued to subside. The basin is bounded by several faults, including the South Margin of Qinghai Lake-Baoji, South Margin of Qinghai Nanshan, Wayuxiangka-Guinan and Wahongshan-Wenquan faults (Tapponnier & Molnar 1977; Hetzel 2013). The basin's sediments comprise mainly Quaternary alluviums and fluvial–lacustrine accumulations as well as Palaeoproterozoic–Neogene lacustrine units (Craddock et al. 2011, 2014; Lu et al. 2012). The basement of the basin is composed of Triassic metamorphic rocks and intrusive granites (Weinert et al. 2020, 2021).

Regional tectonic and geological map of the Gonghe Basin. The inset in the lower left-hand corner shows the regional tectonic around the Gonghe Basin. The red dashed box is the Qabqa geothermal area. The grey dots DR9, DR6, DR4, DR3, GR1 and GR2 are the HDR boreholes. The beach ball indicates the epicentre location and focal mechanism solution of the Mw 6.4 earthquake that occurred in the Gonghe Basin on 26 April 1990. Red lines are basin-bounding faults.
Figure 1.

Regional tectonic and geological map of the Gonghe Basin. The inset in the lower left-hand corner shows the regional tectonic around the Gonghe Basin. The red dashed box is the Qabqa geothermal area. The grey dots DR9, DR6, DR4, DR3, GR1 and GR2 are the HDR boreholes. The beach ball indicates the epicentre location and focal mechanism solution of the Mw 6.4 earthquake that occurred in the Gonghe Basin on 26 April 1990. Red lines are basin-bounding faults.

The Gonghe Basin is renowned for its abundant reservoir of hot dry rock (HDR) geothermal resources (Feng et al. 2018; Gao et al. 2018). It is estimated that approximately 3000 km2 of the basin is enriched with HDR. Notably, within the Qabqa geothermal area, six HDR boreholes, namely DR9, DR6, DR4, DR3, GR1 and GR2, have been drilled (Fig. 1). The average geothermal gradient in HDR borehole GR1 for example, reached 8.8 °C/100 m below 3366 m, whereas the temperature at the bottom of the borehole (3705 m) was high at 236 °C (Zhang et al. 2018; Zhang et al. 2020; Lin et al. 2021). The Gonghe Basin is the most promising HDR development area in China today. Consequently, exploring the spatial distribution characteristics of granites, subsurface temperatures and deep heat sources is crucial for estimating the total HDR resources and assessing their commercial value.

During the subsidence of the sedimentary basin, the sediments undergo progressive compaction and dehydration due to pressure, forming a denser, low-permeability barrier layer. When the burial depth of the mudstone is less than 1500 m, it exhibits lower diagenetic maturity and higher porosity, resulting in weaker sealing capability. Optimal sealing capability is obtained in the depth range of 1500–3200 m. However, when burial depth exceeds 3200 m, the mudstone undergoes higher diagenetic maturity, increased brittleness and increased fracturing under elevated formation pressures, subsequently reducing sealing capability. This layer serves as an effective shield against the migration and diffusion of deep-seated hydrocarbons, thus promoting the accumulation and concentration of basin oil and gas (e.g. Ni et al. 2014a). Additionally, the sedimentary layer acts as a thermal barrier, effectively insulating and preserving the deep-seated heat flow, with important subsequent implications for geothermal resource exploration and development (e.g. Chen et al. 2000; Fuchs 2018; Schütz et al. 2018; Wang et al. 2018; Gascuel et al. 2020; Li et al. 2020). Therefore, accurately determining the depth extent of the sediments in the Gonghe Basin is vital for exploring and developing HDR resources.

Moreover, on 26 April 1990, the Gonghe Basin experienced a Mw 6.4 earthquake (Xu & Chen 1997; Hao et al. 2012; Xie et al. 2018). As seismic waves propagate through the sedimentary basin, they undergo multiple reflections and refractions that resonate with the ultra-soft sediment. This resonance leads to significant ground motion and potentially severe structural damage to buildings (e.g. Chen & Nábelek 1988; Taborda & Bielak 2013; Isbiliroglu et al. 2015; O'Kane & Copley 2020). Consequently, acquiring a comprehensive understanding of the sediment structure is important for assessing seismic hazards in the Gonghe Basin.

Numerous passive seismic methods can be used to constrain the depth and wave speed structure of sedimentary basins. These techniques include ambient noise tomography (e.g. Shapiro et al. 2005; Huang et al. 2010; Lin et al. 2013; Li et al. 2016; Jia & Clayton 2021), horizontal-to-vertical spectral ratio (HVSR) analysis (e.g. Civico et al. 2017) and analysis of converted waves from local earthquakes (e.g. Hough 1990; Li et al. 2014; Ni et al. 2014b; Bao et al. 2018, 2019). P-to-S converted waves (Ps wave) and S-to-P converted waves (Sp wave) from the sedimentary basement and teleseismic receiver function analysis can also be used to investigate the sediment structures (e.g. Hough 1990; Li et al. 2014; Ni et al. 2014b; Yu et al. 2015; Bao & Niu 2017; Jiang et al. 2021). For example, using Ps wave and Sp wave waveform fitting and Sp/S spectral ratio methods, Chen et al. (1996), Langston (2003) and Chiu et al. (2016) studied the thickness and S-wave velocity of sediment in the Mississippi embayment basin, USA.

In this study, we examine the seismograms of the teleseismic earthquakes to identify Ps converted wave energy from the basement of the basin. The delay time between the first-arriving P wave and the Ps conversion is determined via cross-correlation analysis. Subsequently, we establish a correlation between the delay time and the sediment thickness by using the empirical formula of wave velocity variation with depth. Through this process, we obtain the sediment thickness in the Qabqa geothermal area within the Gonghe basin. To validate our results, we compare them with those obtained using the teleseismic receiver function analysis. We use the sediment thickness model to simulate strong ground motion and assess seismic hazards. Additionally, we use a synthetic modelling approach to examine the relationship between the delay time and the sediment parameters and structure, including the seismic ray parameter, sediment thickness, velocity and inclination. We also thoroughly discuss the factors that influence the accuracy of sediment thickness estimation using the delay time approach.

2 DATA AND PRE-PROCESSING

In this study, we utilized data from a short-period dense seismic array of 88 three-component portable seismometers (resonant frequency 4.5 Hz) with a frequency band range of 0.2–100 Hz. These seismometers were deployed in the Qabqa geothermal area in the Gonghe Basin, Qinghai Province, China (Fig. 1). To ensure high-quality imaging of sedimentary structures in the study area, the array was designed to cover the east-central part of the Gonghe Basin with an interstation spacing of approximately 3 km (Fig. 2a). During the deployment of the array (6 August to 13 September 2018), 51 earthquakes of Mw > 5.0 in the epicentral distance range of 30–95° were recorded (Fig. 2b). This data set of high-quality teleseismic waveforms serves as the foundation of our investigation into the sedimentary structures of the Qabqa geothermal area.

(a) Distribution of stations of the short-period dense seismic array in the Qabqa geothermal area in the Gonghe Basin. (b) Distribution of seismic events (Mw > 5.0) during the period (6 August to 13 September 2018).
Figure 2.

(a) Distribution of stations of the short-period dense seismic array in the Qabqa geothermal area in the Gonghe Basin. (b) Distribution of seismic events (Mw > 5.0) during the period (6 August to 13 September 2018).

To prepare the raw teleseismic waveforms for further analysis, a series of pre-processing steps were performed. After instrument response removal, seismograms were demeaned, detrended and bandpass filtered with corner frequencies of 0.2 and 5 Hz. The bandpass filter range was determined via examination of the spectral characteristics of the seismograms (Fig. S1). Seismograms were then transformed into the L-Q-T (P, SV and SH) coordinate system following the approach of Kennett (1991) (Fig. 3).

Example of comparison between the two stations' original and pre-processed waveforms. (a) Original Z-N-E three-component waveforms of station GH817. (b) Pre-processed L-Q-T three-component waveforms of GH817. (c) Original Z-N-E three-component waveforms of GH887. (d) Pre-processed L-Q-T three-component waveforms of GH887.
Figure 3.

Example of comparison between the two stations' original and pre-processed waveforms. (a) Original Z-N-E three-component waveforms of station GH817. (b) Pre-processed L-Q-T three-component waveforms of GH817. (c) Original Z-N-E three-component waveforms of GH887. (d) Pre-processed L-Q-T three-component waveforms of GH887.

3 DISTRIBUTION OF THE DELAY TIME

3.1 Teleseismic Ps-wave method

When a seismic P-wave propagates through an interface between the sediment and the crystalline basement at an oblique angle, it undergoes partial transmission as a P wave (the direct P wave) and partial conversion to an S wave (P-to-S converted wave). Because the sediment velocity is relatively low and the teleseismic P wave typically has a small incidence angle, most P-wave energy is recorded in the L-component. In contrast, the Ps-wave energy, as the S wave, is recorded in the Q-component. By analysing SV P-wave receiver function (SV PRF) waveforms, the coherent Ps wave can be identified (Fig. 4).

Schematic diagram of the propagation path of the teleseismic P wave after it is incident on the sediment. L and Q are seismogram components in direction of P wave propagation and aligned in the direction of the SV phase movement components waveforms of the structure impulse response recorded by the station, respectively. SV PRF is P-wave receiver function, which is calculated by deconvolving the Q from L component waveforms.
Figure 4.

Schematic diagram of the propagation path of the teleseismic P wave after it is incident on the sediment. L and Q are seismogram components in direction of P wave propagation and aligned in the direction of the SV phase movement components waveforms of the structure impulse response recorded by the station, respectively. SV PRF is P-wave receiver function, which is calculated by deconvolving the Q from L component waveforms.

In the absence of seismic anisotropy, the delay time between the direct P wave and the later-arriving basement Ps conversion will depend on the thickness and velocity structure of the basin sediments and the ray parameter. For a horizontally layered medium, the delay time between the P waves and the Ps waves can be given as (Zhu & Kanamori 2000):

(1)

where |${\rm{\Delta }}{t_{Ps}}( {{\rm{\Delta }}{t_{Ps}} = {t_{Ps}} - {t_P}} )$| is the delay time between the P waves and the Ps waves. h is the sediment thickness. |${V_P}$| and |${V_S}$|⁠, respectively represents the average velocity of P and S waves in the sediments. p is the seismic wave ray parameter. Given the small incidence angle of the P waves, it can be treated as nearly vertical incidence. Consequently, the contribution of p can be neglected, resulting in an approximation of |${\rm{\Delta }}{t_{Ps}}$| as:

(2)

Therefore, after obtaining |${\rm{\Delta }}{t_{Ps}}$|⁠, given |${V_P}$| and |${V_S}$| of the study area, h of the study area can be derived as:

(3)

3.2 The delay time |$\Delta {t_{Ps}}$| from the waveform cross-correlation method

3.2.1 Methodology

As the P waves and the Ps waves are both generated by the incident P wave, they should be identical in waveform in ideal conditions. Nevertheless, due to the high-attenuation nature of the sediment, some waveform changes may occur. Despite this, the waveforms of the P and Ps waves still exhibit significant similarities. Consequently, the waveform cross-correlation method can be used to accurately determine the delay time |$\Delta {t_{Ps}}$| between the P waves and the Ps waves.

After selecting teleseismic waveforms with a high signal-to-noise ratio (Table S1) and manually picking first arrivals, we can estimate the delay time |$\Delta {t_{Ps}}$| from the maximum cross-correlation coefficient. The correlation functions of the L-component and Q-component can be calculated by:

(4)

where |${\rm Cor}( \tau )$| is normalized cross-correlation coefficient in the range of [−1, 1]. |${\widehat {{\rm Cor}}_{QL}}( \tau )$| is the correlation functions of the L-component (⁠|$L( t )$|⁠) and Q-component (⁠|$Q( t )$|⁠). |${\rm{*}}$| denotes the conjugate. |$t,\ N$| are the waveform length and finite sample numbers. |$\tau $| is the sliding time window length.

3.2.2 Synthetic tests

In order to explore the effectiveness of waveform cross-correlation analysis in distinguishing between thin and thick sedimentary layers, we devised two synthetic models. Both models comprise two horizontal layers: an upper sedimentary layer atop a homogeneous half-space. The first model comprises a relatively thin sedimentary layer (⁠|$h = 0.2\ {\rm km},\ {V_P} = 2.36\ {\rm km\,s}^{-1},\ {V_s} = 0.87\ {\rm km\,s}^{-1},\ p = 0.062\ {\rm s\,km}^{-1}$|⁠). The theoretical calculation shows that |$\Delta {t_{Ps}}$| of the synthetic model is 0.14 s. The second model comprises a relatively thick sedimentary layer (⁠|$h = 3.0\ {\rm km},\ {V_P} = 4.04\ {\rm km\,s}^{-1},\ {V_s} = 2.31\ {\rm km\,s}^{-1},\ p = 0.062\ {\rm s\,km}^{-1}$|⁠). The theoretical calculation shows that |$\Delta {t_{Ps}}$| of the synthetic model is 0.57 s.

To determine the optimal frequency band filtering, we conducted a series of experiments applying different bandpass filters to the synthetic seismic waveforms of the L-component and Q-component. Specifically, we used bandpass filters of 0.2–0.5 Hz, 0.2–2 Hz and 0.2–5 Hz, acting simultaneously both on the L and Q components. To accurately measure the delay time of Ps waves lagging behind P waves, we calculated the cross-correlation coefficient with |$\tau = - 2\sim 6s,{\rm{\ \ }}t = 1.5s$|⁠. Our results demonstrate that the waveform cross-correlation method is best suited for the 0.2–5 Hz frequency range, offering excellent resolution for both thin and thick layers (Figs 56).

Thin sedimentary layer synthetic tests for the waveform cross-correlation method. (a), (d) and (g) are the synthetic seismic waveforms of the L-component and Q-component with the band-pass filters of 0.2–0.5 Hz, 0.2–2 Hz and 0.2–5 Hz, respectively. (b), (e) and (h) are the processes of calculating the cross-correlation coefficients of the L-component and Q-component by sliding Q-component. (c), (f) and (g) are the results of waveform cross-correlation.
Figure 5.

Thin sedimentary layer synthetic tests for the waveform cross-correlation method. (a), (d) and (g) are the synthetic seismic waveforms of the L-component and Q-component with the band-pass filters of 0.2–0.5 Hz, 0.2–2 Hz and 0.2–5 Hz, respectively. (b), (e) and (h) are the processes of calculating the cross-correlation coefficients of the L-component and Q-component by sliding Q-component. (c), (f) and (g) are the results of waveform cross-correlation.

Thick sedimentary layer synthetic tests for the waveform cross-correlation method. (a), (d) and (g) are the synthetic seismic waveforms of the L-component and Q-component with the bandpass filters of 0.2–0.5 Hz, 0.2–2 Hz and 0.2–5 Hz, respectively. (b), (e) and (h) are the processes of calculating the cross-correlation coefficients of the L-component and Q-component by sliding Q-component. (c), (f) and (g) are the results of waveform cross-correlation.
Figure 6.

Thick sedimentary layer synthetic tests for the waveform cross-correlation method. (a), (d) and (g) are the synthetic seismic waveforms of the L-component and Q-component with the bandpass filters of 0.2–0.5 Hz, 0.2–2 Hz and 0.2–5 Hz, respectively. (b), (e) and (h) are the processes of calculating the cross-correlation coefficients of the L-component and Q-component by sliding Q-component. (c), (f) and (g) are the results of waveform cross-correlation.

3.2.3 Results

After evaluating the preprocessed waveforms and determining the theoretical delay time |$\Delta {t_{Ps}}{\boldsymbol{\ }}$| under geological conditions, we estimated the cross-correlation coefficients with |$\tau = - 0.3\sim 6{\rm{\ }}s,{\rm{\ \ }}t = 1.5{\rm{\ }}s$|⁠. Through a rigorous selection of the waveform cross-correlation results, we eventually selected 64 stations’ Ps and P waves to obtain the spatial distribution of the delay time |$\Delta {t_{Ps}}$| of the dense seismic array in the Gonghe Basin (Fig. 7). The delay time |$\Delta {t_{Ps}}$| gradually increases from east to west and ranges from 0.36 to 1.02 s. For most stations, the standard deviation of the delay time |$\Delta {t_{Ps}}$| is less than 0.25 s.

Distribution of the delay time ${\rm{\Delta }}{t_{Ps}}$ from the waveform cross-correlation analysis of the short-period dense seismic array in the Gonghe Basin. (a) The counts of the Ps waveform were used for each station. (b) The standard deviation of the delay time for each station.
Figure 7.

Distribution of the delay time |${\rm{\Delta }}{t_{Ps}}$| from the waveform cross-correlation analysis of the short-period dense seismic array in the Gonghe Basin. (a) The counts of the Ps waveform were used for each station. (b) The standard deviation of the delay time for each station.

3.3 The delay time |$\Delta {t_{Ps}}$| from the teleseismic receiver function method

3.3.1 Methodology

Receiver functions—time-series, computed from three-component seismograms—record the relative response of Earth structure below a seismograph station. The presence of the sedimentary layer affects P-wave receiver functions, causing the initial P-wave peak to partially represent the Ps wave. This initial peak exhibits a significant time delay compared to the P wave arriving at zero time, and this delay increases with sediment thickness. Therefore, by measuring the arrival time of the initial wave peak of the P-wave receiver functions, we can calculate the sediment thickness (e.g. Zheng et al. 2005; Borah et al. 2015; Agostinetti et al. 2018; Liu et al. 2018; Cunningham & Lekic 2020; Li et al. 2021).

P-wave receiver functions are obtained by deconvolving the waveforms of the Q-component by the L-component. In this study, we utilized the iterative time-domain deconvolution method of Ligorria & Ammon (1999) to calculate the P-wave receiver functions. Additionally, we applied a low-pass Gaussian filter to the receiver functions to remove high-frequency noise. The Gaussian filter in the frequency domain can be expressed as,

(5)

where |$\omega $| is the angular frequency. |$\alpha $| is the Gaussian factor, which represents the width of the Gaussian.

To assess the effectiveness of the receiver function method for estimating sediment thickness, we used the same models as in the earlier trials. We used a Gaussian factor of 10 for these simulations. According to the results of our analysis, sedimentary layers of varying thickness, including both thin and thick layers, could be resolved very well using the filtering frequency band of 0.2–5 Hz (Figs 89).

Thin sedimentary layer synthetic tests for the receiver function method. (a), (d) and (c) are the synthetic seismic waveforms of the L-component and Q-component and receiver function with the band-pass filters of 0.2–0.5 Hz, 0.2–2 Hz and 0.2–5 Hz, respectively.
Figure 8.

Thin sedimentary layer synthetic tests for the receiver function method. (a), (d) and (c) are the synthetic seismic waveforms of the L-component and Q-component and receiver function with the band-pass filters of 0.2–0.5 Hz, 0.2–2 Hz and 0.2–5 Hz, respectively.

Thick sedimentary layer synthetic tests for the receiver function method. (a), (d) and (c) are the synthetic seismic waveforms of the L-component and Q-component and receiver function with the band-pass filters of 0.2–0.5 Hz, 0.2–2 Hz and 0.2–5 Hz, respectively.
Figure 9.

Thick sedimentary layer synthetic tests for the receiver function method. (a), (d) and (c) are the synthetic seismic waveforms of the L-component and Q-component and receiver function with the band-pass filters of 0.2–0.5 Hz, 0.2–2 Hz and 0.2–5 Hz, respectively.

3.3.2 Parameter tests

A Gaussian filter is applied to the receiver functions to remove high-frequency noise. As the width of the Gaussian factor is increased, more noise remains in the receiver functions. Therefore, there is a trade-off between the vertical resolution of the true strata and noise signal artefacts when imaging subsurface structures (Ward & Lin 2017). The ideal Gaussian factor depends on the data and research goals. We established that a Gaussian factor of 5.0 was the ideal compromise between removing high-frequency noise and guaranteeing that the Ps and PpPs phases did not overlap by examining seven alternative Gaussian factors ranging from 1.0 to 10.0 (maximum cut-off frequencies 0.50–5.0 Hz; Figs 1011).

The Ps-wave arrival times of each station obtained from the P-wave receiver functions with Gaussian factors of 1.0–10.0 for a teleseismic event.
Figure 10.

The Ps-wave arrival times of each station obtained from the P-wave receiver functions with Gaussian factors of 1.0–10.0 for a teleseismic event.

P-wave receiver function waveforms with Gaussian factors of 1.0–10.0 of a teleseismic event of two stations. (a) P-wave receiver function waveforms of station GH817. (b) P-wave receiver function waveforms of GH887.
Figure 11.

P-wave receiver function waveforms with Gaussian factors of 1.0–10.0 of a teleseismic event of two stations. (a) P-wave receiver function waveforms of station GH817. (b) P-wave receiver function waveforms of GH887.

3.3.3 Results

We calculated the P-wave receiver functions for each station using the iterative time-domain deconvolution method with a Gaussian factor of 5.0. These receiver functions were then stacked linearly. After quality control, we selected 62 stations’ stacked receiver functions to estimate the Ps-wave arrival times from the initial wave peak with a maximum search range of 1.5 s. The calculated delay time |$\Delta {t_{Ps}}$| ranges from 0.30 to 1.04 s and increases from east to west. For most stations, the standard deviation of the delay time |$\Delta {t_{Ps}}$| is less than 0.25 s (Fig. 12).

Distribution of the delay time ${\rm{\Delta }}{t_{Ps}}$ from the teleseismic receiver function method of the short-period dense seismic array in the Gonghe Basin. (a) The counts of the Ps waveform were used for each station. (b) The standard deviation of the delay time for each station.
Figure 12.

Distribution of the delay time |${\rm{\Delta }}{t_{Ps}}$| from the teleseismic receiver function method of the short-period dense seismic array in the Gonghe Basin. (a) The counts of the Ps waveform were used for each station. (b) The standard deviation of the delay time for each station.

4 RESULTS AND DISCUSSION

4.1 Sediment thickness results

HDR borehole data from the Gonghe basin has revealed the Qabqa geothermal area's two-layered geological structure. The shallow layer comprises Cenozoic sediments, while the deep layer comprises Middle and Late Triassic granites. HDR boreholes DR9, DR6, DR4, DR3, GR1 and GR2 have sediment thicknesses of 1600, 1700, 1340, 1402, 1350 and 940 m. The core geological data of borehole GR1 reveals that the shallow sedimentary layer is a ∼ 500-m-thick Quaternary stratum of sand, gravel and siltstone. And the bottom half of the sedimentary layer is an ∼850-m-thick Palaeoproterozoic–Neoproterozoic layer composed primarily of mudstone and sandy conglomerate (Zhang et al. 2020). Based on the lithological data of the basin, we assigned the average velocity of |${V_s} = 1.40\ {\rm km\,s}^{-1}$| and |${V_p} = 3.80\ {\rm km\,s}^{-1}$| for the sediment layer (Jia et al. 2019; Wang et al. 2019). Using eq. (3), we calculated the distribution of sediment thickness beneath the short-period dense seismic array in the Gonghe Basin (Figs 14a and b).

Since wave velocity varies with depth, sediments with different thicknesses have varied average wave velocities. Thicker sediments tend to have higher average wave velocities than thinner ones (Brocher 2008). We built a new wave velocity model based on an empirical equation that characterizes the change in wave velocity with depth to integrate this variability (Table 1, Fig. 13a). By applying published connections between wave velocity and sediment thickness (Brocher 2005, 2008; Bao et al. 2018), we developed the association between the delay time and sediment thickness (Fig. 13b). We were able to convert delay time to sediment thickness:

(6)

where h is the sediment thickness, and |${\rm{\Delta }}{t_{Ps}}$| is the delay time between the Ps and P waves.

Time–depth conversion relationship. (a) Relationship of ${V_p}$ and ${V_s}$ versus depth. (b) Relationship of delay time and sediment thickness.
Figure 13.

Time–depth conversion relationship. (a) Relationship of |${V_p}$| and |${V_s}$| versus depth. (b) Relationship of delay time and sediment thickness.

Table 1.

Relationship of |${V_p}$| and |${V_s}$| versus depth.

RelationDepth rangeRock type
|${V_p} = 0.7 + 42.698h - 575.8{h^2} + 2931.6{h^3} - 3977.6{h^4}$||$( {h < 0.04\ {\rm km}} )$|Plio-Quaternary deposits
|${V_s} = 0.215 + 10.932h - 138.1{h^2}$||$( {h < 0.04\ {\rm km}} )$|Plio-Quaternary deposits
|${V_p} = 1.5 + 3.735h - 3.543{h^2}$||$( {0.04\ {\rm km} < h < 0.5\ {\rm km}} )$|Late-Pleistocene deposits
|${V_p} = 2.24 + 0.6h$||$( {0\sim 4\ {\rm km}} )$|Tertiary sed. rocks
|${V_s} = 1.5{h^{\frac{1}{4}}}$||$( {0\sim 4\ {\rm km}} )$|Tertiary sed. rocks
|${V_p} = 6.20 + 0.002( {h - 4} )$|
|${V_s} = 2.88 + 0.52( {{V_p} - 5.25} )$|
|$( {h > 4\ {\rm km}} )$|
|$( {h > 4\ {\rm km}} )$|
Tertiary granites
Tertiary granites
RelationDepth rangeRock type
|${V_p} = 0.7 + 42.698h - 575.8{h^2} + 2931.6{h^3} - 3977.6{h^4}$||$( {h < 0.04\ {\rm km}} )$|Plio-Quaternary deposits
|${V_s} = 0.215 + 10.932h - 138.1{h^2}$||$( {h < 0.04\ {\rm km}} )$|Plio-Quaternary deposits
|${V_p} = 1.5 + 3.735h - 3.543{h^2}$||$( {0.04\ {\rm km} < h < 0.5\ {\rm km}} )$|Late-Pleistocene deposits
|${V_p} = 2.24 + 0.6h$||$( {0\sim 4\ {\rm km}} )$|Tertiary sed. rocks
|${V_s} = 1.5{h^{\frac{1}{4}}}$||$( {0\sim 4\ {\rm km}} )$|Tertiary sed. rocks
|${V_p} = 6.20 + 0.002( {h - 4} )$|
|${V_s} = 2.88 + 0.52( {{V_p} - 5.25} )$|
|$( {h > 4\ {\rm km}} )$|
|$( {h > 4\ {\rm km}} )$|
Tertiary granites
Tertiary granites

*|${V_p}$| and |${V_s}$| in |${\rm km\,s}^{-1}$|⁠; depth, h, in km (Brocher 2005, 2008; Bao et al. 2018).

Table 1.

Relationship of |${V_p}$| and |${V_s}$| versus depth.

RelationDepth rangeRock type
|${V_p} = 0.7 + 42.698h - 575.8{h^2} + 2931.6{h^3} - 3977.6{h^4}$||$( {h < 0.04\ {\rm km}} )$|Plio-Quaternary deposits
|${V_s} = 0.215 + 10.932h - 138.1{h^2}$||$( {h < 0.04\ {\rm km}} )$|Plio-Quaternary deposits
|${V_p} = 1.5 + 3.735h - 3.543{h^2}$||$( {0.04\ {\rm km} < h < 0.5\ {\rm km}} )$|Late-Pleistocene deposits
|${V_p} = 2.24 + 0.6h$||$( {0\sim 4\ {\rm km}} )$|Tertiary sed. rocks
|${V_s} = 1.5{h^{\frac{1}{4}}}$||$( {0\sim 4\ {\rm km}} )$|Tertiary sed. rocks
|${V_p} = 6.20 + 0.002( {h - 4} )$|
|${V_s} = 2.88 + 0.52( {{V_p} - 5.25} )$|
|$( {h > 4\ {\rm km}} )$|
|$( {h > 4\ {\rm km}} )$|
Tertiary granites
Tertiary granites
RelationDepth rangeRock type
|${V_p} = 0.7 + 42.698h - 575.8{h^2} + 2931.6{h^3} - 3977.6{h^4}$||$( {h < 0.04\ {\rm km}} )$|Plio-Quaternary deposits
|${V_s} = 0.215 + 10.932h - 138.1{h^2}$||$( {h < 0.04\ {\rm km}} )$|Plio-Quaternary deposits
|${V_p} = 1.5 + 3.735h - 3.543{h^2}$||$( {0.04\ {\rm km} < h < 0.5\ {\rm km}} )$|Late-Pleistocene deposits
|${V_p} = 2.24 + 0.6h$||$( {0\sim 4\ {\rm km}} )$|Tertiary sed. rocks
|${V_s} = 1.5{h^{\frac{1}{4}}}$||$( {0\sim 4\ {\rm km}} )$|Tertiary sed. rocks
|${V_p} = 6.20 + 0.002( {h - 4} )$|
|${V_s} = 2.88 + 0.52( {{V_p} - 5.25} )$|
|$( {h > 4\ {\rm km}} )$|
|$( {h > 4\ {\rm km}} )$|
Tertiary granites
Tertiary granites

*|${V_p}$| and |${V_s}$| in |${\rm km\,s}^{-1}$|⁠; depth, h, in km (Brocher 2005, 2008; Bao et al. 2018).

The study area's sediment thickness distribution range is 500–3000 m after time-depth conversion (Figs 14c and d). Interestingly, this matches HDR borehole data (Table 2). Sediment thickness correlated with the research area's surface topography, increasing from east to west. The eastern region, located in the basin's interior, has increased sediment deposition, in contrast to the western region, which is exposed mountain ranges. The sediment structures from the two approaches matched well, with sediment thickness increasing from ∼500 m in the east to ∼3000 m in the west. The thickness discrepancies between the two methods are typically less than 0.5 km, with the maximum thickness reaching 510 m at GH825 (Fig. 15, Table S2).

Sediment thickness results of the Gonghe Basin with different methods. The colours in the circles represent sediment thickness and the sizes of the circles are standard deviations. (a) Waveform cross-correlation method using the average velocity. (b) Teleseismic receiver function method using the average velocity. (c) Waveform cross-correlation method using the velocity variation with depth. (d) Teleseismic receiver function method using the velocity variation with depth.
Figure 14.

Sediment thickness results of the Gonghe Basin with different methods. The colours in the circles represent sediment thickness and the sizes of the circles are standard deviations. (a) Waveform cross-correlation method using the average velocity. (b) Teleseismic receiver function method using the average velocity. (c) Waveform cross-correlation method using the velocity variation with depth. (d) Teleseismic receiver function method using the velocity variation with depth.

Comparison of the results of sediment thicknesses calculated by the waveform cross-correlation method and the teleseismic receiver function method. (a) Sediment thickness differences maps of stations used in this study. (b) A comparison of the sediment thickness from the waveform cross-correlation method and teleseismic receiver function method. (c) A comparison of the sediment thickness from two methods and HDR boreholes data.
Figure 15.

Comparison of the results of sediment thicknesses calculated by the waveform cross-correlation method and the teleseismic receiver function method. (a) Sediment thickness differences maps of stations used in this study. (b) A comparison of the sediment thickness from the waveform cross-correlation method and teleseismic receiver function method. (c) A comparison of the sediment thickness from two methods and HDR boreholes data.

Table 2.

The sediment thickness beneath the short-period dense seismic array.

No.BorecholeLat. (°N)Long. (°E)Thickness (km)Cc thickness (km)Rc thickness (km)
1DR436.303100.6201.341.401.56
2GR136.253100.6301.401.501.68
3GR236.233100.6570.941.491.34
4DR336.246100.6001.351.461.56
5DR636.300100.5801.701.822.07
6DR936.330100.5801.601.891.93
No.BorecholeLat. (°N)Long. (°E)Thickness (km)Cc thickness (km)Rc thickness (km)
1DR436.303100.6201.341.401.56
2GR136.253100.6301.401.501.68
3GR236.233100.6570.941.491.34
4DR336.246100.6001.351.461.56
5DR636.300100.5801.701.822.07
6DR936.330100.5801.601.891.93

Cc, thickness is the result of waveform cross-correlation method. Rc, thickness is the result of receiver function method.

Table 2.

The sediment thickness beneath the short-period dense seismic array.

No.BorecholeLat. (°N)Long. (°E)Thickness (km)Cc thickness (km)Rc thickness (km)
1DR436.303100.6201.341.401.56
2GR136.253100.6301.401.501.68
3GR236.233100.6570.941.491.34
4DR336.246100.6001.351.461.56
5DR636.300100.5801.701.822.07
6DR936.330100.5801.601.891.93
No.BorecholeLat. (°N)Long. (°E)Thickness (km)Cc thickness (km)Rc thickness (km)
1DR436.303100.6201.341.401.56
2GR136.253100.6301.401.501.68
3GR236.233100.6570.941.491.34
4DR336.246100.6001.351.461.56
5DR636.300100.5801.701.822.07
6DR936.330100.5801.601.891.93

Cc, thickness is the result of waveform cross-correlation method. Rc, thickness is the result of receiver function method.

4.2 Factors influencing the accuracy of sediment thickness

The accuracy of obtaining the sediment thickness from Ps waves is affected by the following factors: the error in determining the delay time |${\rm{\Delta }}{t_{Ps}}$|⁠; the confidence with which the P-wave and S-wave speed structure of the sediment is known; and the difference between the actual sediment model and the theoretical model chosen for calculating the sediment thickness.

The determination error of the delay time |${\rm{\Delta }}{t_{Ps}}$| depends largely on the precision of identifying the arrival time of the Ps wave. The waveform cross-correlation method estimates this delay time based on the waveform similarity between P and Ps waves. The delay time is the sliding time window length with the maximum cross-correlation coefficient. Precise P- and Ps-wave arrival times are not required. Due to multiple waves and noise, the sliding time window length may not match the true delay time. Therefore, it is crucial to select an appropriate waveform length that captures the characteristics of both the P wave and Ps wave on the LQ components. This ensures the reliability of the waveform cross-correlation results and reduces the influence of multiple waves on the measurements. Additionally, excluding seismic records with a low signal-to-noise ratio can further minimize errors in determining the delay time |${\rm{\Delta }}{t_{Ps}}$|⁠.

Errors in P- and S-wave velocity values significantly affect the absolute sediment thickness. The wave velocity often increases with depth. In regions with large variability in sediment thickness, relying exclusively on average velocity may result in sediment thickness predictions that are either underestimated or exaggerated. The relative inaccuracy of the interface diminishes as the sediment thickness grows, but the absolute error increases. It is critical to account for the depth-dependent variation in wave velocity in order to achieve appropriate results.

The sediment thickness might be affected by theoretical model inaccuracies. Simplifying multilayer strata with two layers or ignoring interface inclination might cause calculation errors. The magnitude of error resulting from neglecting the multi-layered nature of the medium depends on the epicentral distance. The inaccuracy is insignificant for teleseismic earthquakes with epicentral distance |$\Delta \ge {30^{\circ}} $| (Fig. S2). The degree of depth estimation error caused by mistaking Ps waves from an inclined interface for a horizontal interface increases with interface inclination (Li et al. 2019). When the interface inclination |${\rm dip} \le {10^{\circ} } $|⁠, its effect is generally negligible (Fig. S3). Therefore, the sediment thickness obtained from teleseismic Ps waves can be approximated using eq. (3) for almost horizontal strata.

4.3 Comparison with other geophysical surveys

Many geophysical surveys have shown that the Qabqa geothermal area in the Gonghe Basin has a maximum sediment thickness of ∼3000 m (Wang et al. 2010, 2019 ; Jia et al. 2019). The deep seismic sounding (DSS) P-wave velocity profile (Jia et al. 2019) and receiver function S-wave velocity profile (Wang et al. 2019) show a gradual decrease in P- and S-wave velocity in the Gonghe Basin from east to west, indicating an increase in sediment thickness. This discovery is consistent with the trend of increasing sediment thickness from east to west from the waveform cross-correlation analysis (Figs 16b–e).

(a) The distribution of sediment thickness beneath the short-period dense seismic array in the Gonghe Basin was calculated by the waveform correlation method. (b) P-wave velocity profile of deep seismic sounding (DSS) in the Gonghe Basin (modified from Jia et al. 2019). (c) S-wave velocity profile of receiver function in the Gonghe Basin (modified from Wang et al. 2019). (d) Sediment thickness distribution beneath the profile AA'. (e) Sediment thickness distribution beneath the profile BB'. (f) Sediment thickness distribution beneath the profile CC'. (g) Sediment thickness distribution beneath the profile DD'. (f) Sediment thickness distribution beneath the profile EE'.
Figure 16.

(a) The distribution of sediment thickness beneath the short-period dense seismic array in the Gonghe Basin was calculated by the waveform correlation method. (b) P-wave velocity profile of deep seismic sounding (DSS) in the Gonghe Basin (modified from Jia et al. 2019). (c) S-wave velocity profile of receiver function in the Gonghe Basin (modified from Wang et al. 2019). (d) Sediment thickness distribution beneath the profile AA'. (e) Sediment thickness distribution beneath the profile BB'. (f) Sediment thickness distribution beneath the profile CC'. (g) Sediment thickness distribution beneath the profile DD'. (f) Sediment thickness distribution beneath the profile EE'.

Sediment thickness distributions were plotted along three profiles (CC', DD' and EE’) crossing the HDR boreholes to compare the result to HDR borehole data (Figs 16f–h). Due to being close to seismic stations, HDR borehole DR4 has the best sediment thickness agreement at 0.06 km. However, seismic stations weakly limit HDR boreholes DR3, GR1, GR2, DR6 and DR9, resulting in larger discrepancies. These boreholes have thickness differences of 0.16, 0.10, −0.55, 0.12 and −0.29 km.

4.4 Strong ground motion simulation

In thick sedimentary layers, a significant impedance contrast between the sediment and bedrock causes a low fundamental frequency and a high amplitude peak, amplifying ground motions. Large structures, including bridges, high-rise buildings and industrial facilities, are at risk from low-frequency ground shaking (Cunningham & Lekic 2020). The seismic ambient noise horizontal-to-vertical spectral ratio (HVSR) is usually used to identify sediment layers’ fundamental frequency and amplitude peak, indicating the S-wave resonant frequency. We used the procedures given by Bao et al. (2019) to calculate the Qabqa geothermal area's HVSR (Figs 17b and c). There is a definite relationship between the sediment thickness and the fundamental frequency as well as the amplitude peak. Specifically, as the sediment layer thickness grows, the fundamental frequency lowers (Fig. 18a) while the amplitude peak increases (Fig. 18b).

(a) The distribution of sediment thickness beneath the short-period dense seismic array in the Gonghe Basin was calculated by the waveform correlation method. (b) The fundamental frequency and (c) amplitude peak of the sediment thickness model were calculated by the HVSR method. (d) Peak ground velocity (PGV) in the Gonghe Basin considering the sediment thickness model and topography.
Figure 17.

(a) The distribution of sediment thickness beneath the short-period dense seismic array in the Gonghe Basin was calculated by the waveform correlation method. (b) The fundamental frequency and (c) amplitude peak of the sediment thickness model were calculated by the HVSR method. (d) Peak ground velocity (PGV) in the Gonghe Basin considering the sediment thickness model and topography.

(a) The fundamental frequency decreases with the sediment thickness. (b) The amplitude peak increases with the sediment thickness. (c) The PGV increases with the sediment thickness.
Figure 18.

(a) The fundamental frequency decreases with the sediment thickness. (b) The amplitude peak increases with the sediment thickness. (c) The PGV increases with the sediment thickness.

Ground motion in sedimentary basins is complicated by intrabasin surface waves, basin edge effects and basin focusing effects. We simulated strong ground motion from the 1990 Mw 6.4 earthquake using our Gonghe Basin sediment thickness model to examine the significance of the sediment thickness model for seismic hazard assessment. The 3-D seismic wave curve grid finite difference method was used for the simulation (Zhang et al. 2012; Sun et al. 2018). In the calculations, surface topography data from the GTOPO30 DEM was used. Based on the depth-dependent seismic wave velocity model (Table 1), we modeled the earthquake as a point source at 32 km deep and simulated low-frequency strong ground motion (1 Hz and below). Our simulation results show that asymmetric PGV distributions extend to the west and northwest from the earthquake epicentre (Fig. 17d). The PGV is amplified by thicker sediments in the northwest (Fig. 18c). Therefore, our sediment thickness model provides valuable insights for ground motion simulation and seismic hazard quantification in the Gonghe Basin.

6 CONCLUSIONS

In this study, we deployed a dense seismic array of 88 short-period three-component seismometers in the Gonghe Basin in the northeast Tibetan Plateau. Using teleseismic waveforms recorded by the array, we extracted the delay time between the Ps waves from the sediment basement and the P waves by waveform cross-correlation analysis. Based on the delay time, empirical sediment velocity structure model, and HDR borehole data, we inferred a reliable sediment thickness model for the Qabqa geothermal area in the basin. Additionally, we compared this model with those derived from teleseismic receiver function analysis. We find that the sediment thickness of the Qabqa geothermal area of the Gonghe Basin gradually increases from ∼500 m in the east to ∼3000 m in the west. The thick sediment layers provide an ideal thermal storage cover for HDR. Furthermore, our strong ground motion simulations utilizing the sediment thickness model revealed that thick sedimentary layers may amplify seismic waves, causing significant seismic hazards.

Through theoretical analysis and data testing, we find that waveform cross-correlation analysis of teleseismic earthquakes recorded at the short-period dense seismic array can constrain the delay time between the Ps and P waves to construct sediment thickness. During the one-month observation period, the array can capture at least one or two teleseismic earthquakes of Mw > 5 per day with high signal-to-noise ratios. This is sufficient to detect sediment structure without relying on local earthquakes, which are typically difficult to gather due to time and space constraints. Thus, by deploying a dense array in the targeted sedimentary basin area, direct identification of teleseismic converted waves from the sedimentary basement is possible, providing valuable insights into the sediment structure.

SUPPORTING INFORMATION

Please note: Oxford University Press is not responsible for the content or functionality of any supporting materials supplied by the authors. Any queries (other than missing material) should be directed to the corresponding author for the paper.

ACKNOWLEDGEMENTS

We thank the editor and reviewers for their valuable comments that improve the paper. All figures in this study are plotted by the Generic Mapping Tools version 6 (Wessel et al. 2019). The GTOPO30 DEM from United States Geological Survey (https://www.usgs.gov/centers/eros/science/usgs-eros-archive-digital-elevation-global-30-arc-second-elevation-gtopo30). This work is supported by the National Natural Science Foundation of China (41974064 and 42174076) and China Geological Survey (DD20190131).

DATA AVAILABILITY

The data sets are available at: https://doi.org/10.6084/m9.figshare.20479851. The authors declare that there is no conflict of interest regarding the publication of this manuscript.

References

Agostinetti
N.P.
,
Martini
F.
,
Mongan
J.
,
2018
.
Sedimentary basin investigation using receiver function: an East African Rift case study
,
Geophys. J. Int.
,
215
(
3
),
2105
2113
..

Bao
F.
,
Li
Z.
,
Yuen
D.A.
,
Zhao
J.
,
Ren
J.
,
Tian
B.
,
Meng
Q.
,
2018
.
Shallow structure of the Tangshan fault zone unveiled by dense seismic array and horizontal-to-vertical spectral ratio method
,
Phys. Earth planet. Inter.
,
281
,
46
54
..

Bao
F.
,
Li
Z.
,
Tian
B.
,
Wang
L.
,
Tu
G.
,
2019
.
Sediment thickness variations of the Tangshan fault zone in North China from a dense seismic array and microtremor survey
,
J. Asian Earth Sci.
,
185
,
doi:10.1016/j.jseaes.2019.104045
.

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
,
J. geophys. Res.
,
122
(
11
),
9083
9094
..

Borah
K.
,
Kanna
N.
,
Rai
S.S.
,
Prakasam
K.S.
,
2015
.
Sediment thickness beneath the Indo-Gangetic Plain and Siwalik Himalaya inferred from receiver function modelling
,
J. Asian Earth Sci.
,
99
,
41
56
..

Brocher
T.M.
,
2005
.
Empirical relations between elastic wavespeeds and density in the Earth's crust
,
Bull. seism. Soc. Am.
,
95
(
6
),
2081
2092
..

Brocher
T.M.
,
2008
.
Compressional and shear-wave velocity versus depth relations for common rock types in Northern California
,
Bull. seism. Soc. Am.
,
98
(
2
),
950
968
..

Chen
J.
,
Xu
Y.
,
Huang
D.
,
2000
.
Geochemical characteristics and origin of natural gas in Tarim Basin, China
,
AAPG Bull.
,
84
(
5
),
591
606
.

Chen
K.C.
,
Yang
J.M.
,
Tun
Y.
,
1996
.
Shear-wave velocity of the sedimentary basin in the upper Mississippi embayment using S-to-P converted waves
,
Bull. seism. Soc. Am.
,
86
(
3
),
848
856
..

Chen
W.-P.
,
Nábelek
J.
,
1988
.
Seismogenic strike-slip faulting and the development of the North China Basin
,
Tectonics
,
7
(
5
),
975
989
..

Chiu
S.C.C.
,
Langston
C.A.
,
Chiu
J.M.
,
Withers
M.
,
2016
.
Imaging shallow crustal structure in the Upper Mississippi embayment using local earthquake waveform data
,
Bull. seism. Soc. Am.
,
106
(
4
),
1394
1406
..

Civico
R.
et al.
2017
.
Geometry and evolution of a fault-controlled Quaternary basin by means of TDEM and single-station ambient vibration surveys: the example of the 2009 L'Aquila earthquake area, central Italy
,
J. geophys. Res.
,
122
(
3
),
2236
2259
.

Craddock
W.
,
Kirby
E.
,
Zhang
H.
,
2011
.
Late Miocene–Pliocene range growth in the interior of the northeastern Tibetan Plateau
,
Lithosphere
,
3
(
6
),
420
438
..

Craddock
W.H.
,
Kirby
E.
,
Zhang
H.
,
Clark
M.K.
,
Champagnac
J.-D.
,
Yuan
D.
,
2014
.
Rates and style of Cenozoic deformation around the Gonghe Basin, northeastern Tibetan Plateau
,
Geosphere
,
10
(
6
),
1255
1282
..

Cunningham
E.
,
Lekic
V.
,
2020
.
Constraining properties of sedimentary strata using receiver functions: an example from the Atlantic Coastal Plain of the Southeastern United States
,
Bull. seism. Soc. Am.
,
110
(
2
),
519
533
..

Feng
Y.-f.
,
Zhang
X.-x.
,
Zhang
B.
,
Liu
J.-t.
,
Wang
Y.-g.
,
Jia
D.-l.
,
Hao
L.-r.
,
Kong
Z.-y.
,
2018
.
The geothermal formation mechanism in the Gonghe Basin: discussion and analysis from the geological background
,
China Geol.
,
1
(
3
),
331
345
..

Fuchs
S.
,
2018
.
The variability of rock thermal properties in sedimentary basins and the impact on temperature modelling—a Danish example
,
Geothermics
,
76
,
1
14
..

Gao
J.
et al.
2018
.
Three-dimensional magnetotelluric imaging of the geothermal system beneath the Gonghe Basin, Northeast Tibetan Plateau
,
Geothermics
,
76
,
15
25
..

Gascuel
V.
,
Bédard
K.
,
Comeau
F.-A.
,
Raymond
J.
,
Malo
M.
,
2020
.
Geothermal resource assessment of remote sedimentary basins with sparse data: lessons learned from Anticosti Island, Canada
,
Geotherm. Ener.
,
8
(
1
),
doi:10.1186/s40517-020-0156-1
.

Hao
M.
,
Shen
Z.-K.
,
Wang
Q.
,
Cui
D.
,
2012
.
Postseismic deformation mechanisms of the 1990 Mw 6.4 Gonghe, China earthquake constrained using leveling measurements
,
Tectonophysics
,
532-535
,
205
214
..

Hetzel
R.
,
2013
.
Active faulting, mountain growth, and erosion at the margins of the Tibetan Plateau constrained by in situ-produced cosmogenic nuclides
,
Tectonophysics
,
582
,
1
24
..

Hough
S.E.
,
1990
.
Constraining sediment thickness in the San Francisco Bay Area using observed resonances and P-to-S conversions
,
Geophys. Res. Lett.
,
17
(
9
),
1469
1472
..

Huang
Y.-C.
,
Yao
H.
,
Huang
B.-S.
,
van der Hilst
R.D.
,
Wen
K.-L.
,
Huang
W.-G.
,
Chen
C.-H.
,
2010
.
Phase velocity variation at periods of 0.5–3 seconds in the Taipei Basin of Taiwan from correlation of ambient seismic noise
,
Bull. seism. Soc. Am.
,
100
(
5A
),
2250
2263
..

Isbiliroglu
Y.
,
Taborda
R.
,
Bielak
J.
,
2015
.
Coupled soil-structure interaction effects of building clusters during earthquakes
,
Earthq. Spectra
,
31
(
1
),
463
500
..

Jia
S.
,
Guo
W.
,
Mooney
W.D.
,
Wang
F.
,
Duan
Y.
,
Zhao
J.
,
Lin
J.
,
Liu
Z.
,
2019
.
Crustal structure of the middle segment of the Qilian fold belt and the coupling mechanism of its associated basin and range system
,
Tectonophysics
,
770
,
doi:10.1016/j.tecto.2019.06.024
.

Jia
Z.
,
Clayton
R.W.
,
2021
.
Determination of near surface shear-wave velocities in the Central Los Angeles basin with dense arrays
,
J. geophys. Res.
,
126
(
5
),
e2020JB021369
,
doi:10.1029/2020JB021369
.

Jiang
X.
,
Hu
S.
,
Yang
H.
,
2021
.
Depth extent and Vp/Vs ratio of the Chenghai Fault Zone, Yunnan, China constrained from dense-array-based teleseismic receiver functions
,
J. geophys. Res.
,
126
(
8
),
doi:10.1029/2021JB022190
.

Kennett
B.L.N.
,
1991
.
The removal of free surface interactions from three-component seismograms
,
Geophys. J. Int.
,
104
(
1
),
153
154
..

Langston
C.A.
,
2003
.
Local earthquake wave propagation through Mississippi Embayment Sediments, Part I: body-wave phases and local site responses
,
Bull. seism. Soc. Am.
,
93
(
6
),
2664
2684
..

Li
C.
,
Yao
H.
,
Fang
H.
,
Huang
X.
,
Wan
K.
,
Zhang
H.
,
Wang
K.
,
2016
.
3D near-surface shear-wave velocity structure from ambient-noise tomography and borehole data in the Hefei Urban Area, China
,
Seismol. Res. Lett.
,
87
(
4
),
882
892
..

Li
J.
,
Song
X.
,
Wang
P.
,
Zhu
L.
,
2019
.
A generalized H-κ method with harmonic corrections on Ps and its crustal multiples in receiver functions
,
J. geophys. Res.
,
124
,
3782
3801
..

Li
P.
,
Huang
G.-C.D.
,
Savvaidis
A.
,
2021
.
Seismic features of the Permian Basin Region from receiver function analysis
,
Tectonophysics
,
801
,
doi:10.1016/j.tecto.2021.228718
.

Li
X.
,
Cai
L.
,
Liu
S.
,
Li
X.
,
2020
.
Thermal properties of evaporitic rocks and their geothermal effects on the Kuqa Foreland Basin, Northwest China
,
Geothermics
,
88
,
doi:10.1016/j.geothermics.2020.101898
.

Li
Z.
,
Ni
S.
,
Somerville
P.
,
2014
.
Resolving shallow shear-wave velocity structure beneath station CBN by waveform modeling of the Mw 5.8 Mineral, Virginia, earthquake sequence
,
Bull. seism. Soc. Am.
,
104
(
2
),
944
952
..

Ligorria
J.P.
,
Ammon
C.J.
,
1999
.
Iterative deconvolution and receiver-function estimation
,
Bull. seism. Soc. Am.
,
89
(
5
),
1395
1400
..

Lin
F.-C.
,
Li
D.
,
Clayton
R.W.
,
Hollis
D.
,
2013
.
High-resolution 3D shallow crustal structure in Long Beach, California: application of ambient noise tomography on a dense seismic array
,
Geophysics
,
78
(
4
),
Q45
Q56
..

Lin
W.
,
Wang
G.
,
Zhang
S.
,
Zhao
Z.
,
Xing
L.
,
Gzn
H.
,
Tan
X.
,
2021
.
Heat aggregation mechanisms of hot dry rocks resources in the Gonghe Basin, Northeastern Tibetan Plateau
,
Acta Geol. Sin.
,
95
(
6
),
1793
1804
..
(English Edition)

Liu
G.
,
Persaud
P.
,
Clayton
R.W.
,
2018
.
Structure of the Northern Los Angeles basins revealed in teleseismic receiver functions from short-term nodal seismic arrays
,
Seismol. Res. Lett.
,
89
(
5
),
1680
1689
..

Lu
H.
,
Wang
E.
,
Shi
X.
,
Meng
K.
,
2012
.
Cenozoic tectonic evolution of the Elashan range and its surroundings, northern Tibetan Plateau as constrained by paleomagnetism and apatite fission track analyses
,
Tectonophysics
,
580
,
150
161
..

Ni
S.
,
Li
Z.
,
Somerville
P.
,
2014b
.
Estimating subsurface shear velocity with radial to vertical ratio of local P waves
,
Seismol. Res. Lett.
,
85
(
1
),
82
90
..

Ni
Y.
et al.
2014a
.
Studies on gas origin and gas source correlation using stable carbon isotopes—a case study of the giant gas fields in the Sichuan Basin, China
,
Energy Explor. Exploit.
,
32
(
1
),
41
74
..

O'Kane
A.
,
Copley
A.
,
2020
.
The controls on earthquake ground motion in foreland-basin settings: the effects of basin and source geometry
,
Geophys. J. Int.
,
225
(
1
),
512
529
.

Schütz
F.
,
Winterleitner
G.
,
Huenges
E.
,
2018
.
Geothermal exploration in a sedimentary basin: new continuous temperature data and physical rock properties from northern Oman
,
Geotherm. Ener.
,
6
(
1
),
5
,
doi:10.1186/s40517-018-0091-6
.

Shapiro
N.M.
,
Campillo
M.
,
Stehly
L.
,
Ritzwoller
M.H.
,
2005
.
High-resolution surface-wave tomography from ambient seismic noise
,
Science
,
307
(
5715
),
1615
1618
.

Sun
Y.
,
Zhang
W.
,
Chen
X.
,
2018
.
3D seismic wavefield modeling in generally anisotropic media with a topographic free surface by the curvilinear grid finite-difference method
,
Bull. seism. Soc. Am.
,
108
(
3A
),
1287
1301
..

Taborda
R.
,
Bielak
J.
,
2013
.
Ground-motion simulation and validation of the 2008 Chino Hills, California, Earthquake
,
Bull. seism. Soc. Am.
,
103
(
1
),
131
156
..

Tapponnier
P.
,
Molnar
P.
,
1977
.
Active faulting and tectonics in China
,
J. geophys. Res.
,
82
(
20
),
2905
2930
.

Wang
B.
,
He
S.H.
,
Li
B.X.
,
Song
Y.D.
,
2010
.
Geothermal resource distribution characteristics of Gonghe basin in Qinghai-Effect of CSAMT in geothermal prospecting
,
Mineral. Res. Geol.
,
24
,
279
285
.,
(in Chinese)
.

Wang
G.-l.
,
Zhang
W.
,
Ma
F.
,
Lin
W.-j.
,
Liang
J.-y.
,
Zhu
X.
,
2018
.
Overview on hydrothermal and hot dry rock researches in China
,
China Geol.
,
1
(
2
),
273
285
..

Wang
X.
,
Chen
L.
,
Ling
Y.
,
Gao
Y.
,
Zhang
J.
,
Yao
H.
,
2019
.
A new method to constrain shallow crustal S-wave velocities based on direct P-wave amplitudes in receiver functions and its application in northeastern Tibet
,
Sci. China Earth Sci.
,
62
,
1819
1831
.,

Ward
K.M.
,
Lin
F.C.
,
2017
.
On the viability of using autonomous three-component nodal geophones to calculate teleseismic Ps receiver functions with an application to old faithful, Yellowstone
,
Seismol. Res. Lett.
,
88
(
5
),
1268
1278
.,

Weinert
S.
,
Bär
K.
,
Sass
I.
,
2020
.
Thermophysical rock properties of the crystalline Gonghe Basin Complex (Northeastern Qinghai–Tibet-Plateau, China) basement rocks
,
Environ. Earth Sci.
,
79
(
3
),
77
,
doi:10.1007/s12665-020-8808-9
.

Weinert
S.
,
Bär
K.
,
Scheuvens
D.
,
Sass
I.
,
2021
.
Radiogenic heat production of crystalline rocks in the Gonghe Basin Complex (northeastern Qinghai–Tibet plateau, China)
,
Environ. Earth Sci.
,
80
(
7
),
270
,
doi:10.1007/s12665-021-09558-x
.

Wessel
P.
,
Luis
J.F.
,
Uieda
L.
,
Scharroo
R.
,
Wobbe
F.
,
Smith
W.H.F.
,
Tian
D.
,
2019
.
The Generic Mapping Tools version 6
,
Geochem. Geophys. Geosyst.
,
20
,
5556
5564
..

Xie
H.
,
Su
Q.
,
Yuan
D.-y.
,
He
W.-G.
,
Wang
A.-g.
,
Shao
Y.-x.
,
2018
.
Geomorphic evidence of unrecognized Qinghai Nanshan fault in the Gonghe Basin, northeastern Tibetan Plateau
,
Arabian J. Geosci.
,
11
(
18
),
548
.

Xu
L.-S.
,
Chen
Y.-T.
,
1997
.
Source parameters of the Gonghe, Qinghai Province, China, earthquake from inversion of digital broadband waveform data
,
Acta Seismol. Sin.
,
10
(
2
),
143
159
..

Yu
Y.
,
Song
J.
,
Liu
K.H.
,
Gao
S.S.
,
2015
.
Determining crustal structure beneath seismic stations overlying a low-velocity sedimentary layer using receiver functions
,
J. geophys. Res.
,
120
(
5
),
3208
3218
..

Zhang
C.
et al.
2020
.
Radiogenic heat production variations in the Gonghe basin, northeastern Tibetan Plateau: implications for the origin of high-temperature geothermal resources
,
Renew. Ener.
,
148
,
284
297
..

Zhang
C.
,
Jiang
G.
,
Shi
Y.
,
Wang
Z.
,
Wang
Y.
,
Li
S.
,
Jia
X.
,
Hu
S.
,
2018
.
Terrestrial heat flow and crustal thermal structure of the Gonghe-Guide area, northeastern Qinghai-Tibetan plateau
,
Geothermics
,
72
,
182
192
..

Zhang
W.
,
Zhang
Z.
,
Chen
X.
,
2012
.
Three-dimensional elastic wave numerical modelling in the presence of surface topography by a collocated-grid finite-difference method on curvilinear grids
,
Geophys. J. Int.
,
190
(
1
),
358
378
..

Zheng
T.
,
Zhao
L.
,
Chen
L.
,
2005
.
A detailed receiver function image of the sedimentary structure in the Bohai Bay Basin
,
Phys. Earth planet. Inter.
,
152
(
3
),
129
143
..

Zhu
L.
,
Kanamori
H.
,
2000
.
Moho depth variation in southern California from teleseismic receiver functions
,
J. geophys. Res.
,
105
(
B2
),
2969
2980
..

This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://dbpia.nl.go.kr/journals/pages/open_access/funder_policies/chorus/standard_publication_model)

Supplementary data