• <tr id="yyy80"></tr>
  • <sup id="yyy80"></sup>
  • <tfoot id="yyy80"><noscript id="yyy80"></noscript></tfoot>
  • 99热精品在线国产_美女午夜性视频免费_国产精品国产高清国产av_av欧美777_自拍偷自拍亚洲精品老妇_亚洲熟女精品中文字幕_www日本黄色视频网_国产精品野战在线观看 ?

    Teleseismic P-Wave Tomography of the New Guinea-Solomon Arc System

    2022-06-14 04:17:54FENGBingZHANGHaoGONGWeiXINGJunhuiLIDeyongXUChongandANLong
    Journal of Ocean University of China 2022年3期

    FENG Bing, ZHANG Hao, GONG Wei, XING Junhui, LI Deyong, XU Chong, and AN Long

    Teleseismic P-Wave Tomography of the New Guinea-Solomon Arc System

    FENG Bing1), #, ZHANG Hao1), #, GONG Wei1), 2), *, XING Junhui1), 2), LI Deyong1), 2), XU Chong1), and AN Long1)

    1),,,,266100,2),,266237,

    A P-wave tomographic traveltime inversion was applied to obtain a new model of seismic velocity anomalies beneath the New Guinea-Solomon arc system (PN-SL). The P-wave traveltime data, obtained from the revised International Seismological Center catalog, were recorded by 82 seismic stations in the PN-SL. Under the constraints of the epicenter distance, magnitude, and the number of stations recorded, 15009 effective P-wave traveltime data were selected from 2011 teleseismic events. The obtained model showed that the Solomon Sea Plate subducted beneath the New Britain Island along the New Britain Trench at an angle of >70? and that the slab can be traced down to a depth of approximately 800km. Conversely, we cannot observe a high-velocity anomaly exhibited by the subducted Solomon Sea Plate in the deep mantle at the Trobriand Trench, and the slab stopped at a depth of <200km. The double subduction of the Solomon Sea Plate strongly modified the subduction patterns of the early subducted Pacific and Australian plates in the mantle along the West Melanesian Trench and the Pocklington Trough, respectively. In addition, the subducted Solomon Sea Plate induced the melting of the upper mantle to form a low-velocity anomaly, which provided the deep dynamic source for the expansion of the Bismarck Sea. Based on the joint consideration of the tomography results and a petrological analysis, the low-velocity anomalies beneath the Solomon Sea and Woodlark Basin are closely related to the early subduction of the Pacific and Australian Plates, respectively.

    New Guinea-Solomon arc system; traveltime tomography; trench; back-arc basin

    1 Introduction

    The New Guinea-Solomon arc system (PN-SL) exists in the convergent zone between the Pacific Plate and Indo- Australian Plate, surrounded by the Caroline Plate, Australian Plate, and Ontong Java Plateau, which is the largest oceanic plateau on Earth. This complex tectonic and dynamic background causes the PN-SL to exhibit a special structural system characterized by ‘four trenches and three basins’, where the West Melanesian-North Solomon Trench, Bismarck Sea, New Britain-San Cristobal Trench, Solomon Sea, Trobriand Trench, Woodlark Basin, and Pocklington Trough are distributed in a narrow space within a width of less than 1000km. The PN-SL is a region with one of the most active crustal movements and most intense geological processes worldwide (Gong., 2019a). Its unique geographic location and geological structure make it a natural laboratory for subduction dynamic studies. Since the 1970s, scientists have paid considerable attention to this area to reveal the deep structure and tectonic dynamics of the PN-SL (Taylor, 1979; Crook and Taylor, 1994; Taylor., 1995; Mann and Taira, 2004; Wallace., 2004, 2014).

    Since 1900, more than 1000 earthquakes with magnitudes greater than 6.0 have occurred in the PN-SL (https://earthquake.usgs.gov/earthquakes/search/), and this intense seismic activity enabled the determination of the spatial and structural characteristics of the subduction zone. The work of Taylor (1979) provides an overview of the structural characteristics of the Bismarck Sea, including the sea-floor topography, focal mechanisms, and seafloor magneticanomaly data, demarcating the boundary between its sou- thern and northern blocks. The pattern of the Wadati-Beni- off zone gradually changes northward from the double sub- duction on the southern side of the Solomon Islands arc to the northward subduction of the Solomon Sea Plate alongthe New Britain Trench at a high angle (Cooper and Taylor,1985). The natural epicenter and focal mechanism data de- monstrate that the subduction angle of the Solomon Basin along the New Britain Trench reaches 70?, whereas the dip angle of the slab is less than 40? at the Trobriand Trench (Abers and Roecker, 1991; Pegler., 1995). A slab tear exists in the subducted Solomon Sea slab between the New Britain and Solomon Islands, as indicated by the epicenters (Holm and Richards, 2013). Moreover, seismic re- flection data and tomography have been used to explore the structure of the crust and mantle in the PN-SL. The crustal velocities calculated by the deep seismic reflection data present that the Moho surface gradually deepens from the Bismarck Sea toward the New Britain Island (Finlayson., 1972). According to Miura. (2002) and Mann and Taira (2004), the seismic reflection data collected in the North Solomon Trench showed that the crustal thickness of the Ontong Java Plateau can reach 33km. Richardson(2000) presented the Rayleigh-wave tomography of the Ontong Java Plateau using the waveforms of more than 140 earthquakes and observed that the P-wave velocity within the crust-mantle beneath the Ontong Java Plateau is slightly lower than that in the surrounding Paci- fic area. Zelt. (2001) and Abers. (2002) describ- ed the crustal and deep mantle structure of the western side of the Woodlark Basin using the active and passive source seismic data observed by fixed stations and ocean bottom seismometer array, confirming the westward expan- sion of the Woodlark Basin in terms of the deep dynamic characteristics of the crust and mantle. The large-scale seismic tomography study in the Indonesia-Tonga area in the study of Hall and Spakman (2002, 2003) demonstrated the rough and deep geometry of the northward subduction of the Australian Plate along the Pocklington Trough and the southward subduction of the Pacific Plate along the West Melanesian Trench.

    From the above information, although previous studies have explored the subduction geometry of the PN-SL using seismic information, the detailed structure and movement of the slab in the deep mantle still have not been obtained. In addition, the specific relationship between the subducted Solomon Sea Plate and the early subducted Pacific and Australian Plates is still unclear. To counter these problems, we used the recorded P-wave seismic data to analyze the velocity structure of the PN-SL system using the teleseismic tomography method. We used the abundantly available epicenters and P-wave traveltime data to explore the deep subduction geometry of the PN-SL and provide key insights into the subduction dynamics of the PN-SL.

    2 Geological Setting

    The PN-SL is located at the intersection of the Indo-Aus- tralian Plate and Pacific Plate. In the Cenozoic period, the Pacific Plate subducted beneath the Australian Plate along the West Melanesian and North Solomon Trenches (Schellart and Spakman, 2015). Induced by the subduction of the Pacific Plate, the Australian Plate is characterized by a strong tension deformation, forming a series of back-arc basins, such as Pocklington and Coral Sea Basins (Crawford., 2003). In the middle–late Eocene (45Myr), as a new back-arc spreading basin, the Pocklington Basin began to subduct northward under the Pacific Plate. In the late Eocene (40–35Myr), the Solomon Sea Basin expand- ed (Schellart., 2006). During the Oligocene, the Pock- lington Basin completely subducted and disappeared, and the New Guinea Island collided with the passive continental margin on the northern edge of the Australian Plate. Then, the expansion of the Solomon Sea Basin stopped. From the late Oligocene to the early Miocene, the Ontong Java Plateau and Solomon Islands collided and converged, and the subduction movement of the Pacific Plate along the West Melanesian and North Solomon Trenches graduallydecreased (Petterson., 1997). At 8Myr, the subduction activity migrated to the back-arc region, and the subduction of the Solomon Sea Basin along the New Britain Trench initiated (Holm., 2016). At 5Myr, the Woodlark and Bismark Basins expanded, influenced by the sub- duction of the Solomon Sea Basin (., Schellart., 2006; Holm., 2016). Affected by the Cenozoic convergence between these two plates, throughout the critical tectonic periods (50, 45, 25, 8, and 5Myr), the PN-SL gradually formed a complicated subduction system containing a series of subduction zones (Fig.1). The PN-SL includes not only the mature subduction system (over 40 Myr) in the West Melanesian and North Solomon Trenches (Schellart., 2006) but also the subduction system (<8 Myr) of the young Solomon Sea Plate (<35Myr) along the New Britain-San Cristobal and Trobriand Trenches (Holm., 2016).

    The Solomon Sea is a diamond-shaped basin located at the center of the PN-SL and an Eocene-to-early Oligocene back-arc basin (Joshima., 1987; Hall and Spakman, 2003). The average water depth of the Solomon Sea is 4000–5000m. It is one of many microplates in the convergent boundary system of the Indo-Australian and Pacific Plates (Li., 2018). At 25Myr ago, the Ontong Java Plateau collided with the Solomon Islands; this event terminated the southwest-dipping subduction of the Pacific Plate. In addition, the Solomon Sea became bounded by the New Britain Trench with a depth of 6000–8500m at its northern margin and the Trobriand Trench with a depth of 5000m at its southern margin in the late Miocene (Holm., 2016). Compared with the high convergence rate of 9–11cmyr?1at the New Britain Trench, the oblique subduction convergence rate of the Trobriand Trench was slower (2cmyr?1) (Honza., 1989; Tregoning., 1998; Wallace., 2004; DeMets, 2010). With the complete subduction of the Solomon Sea Plate in the western part of the Solomon Sea, the Trobriand and New Britain Trenches gradually merged, resulting in the convergence of New Guinea Island, Finistere Massif, and New Britain Island (Cooper and Taylor, 1987; Pegler., 1995; Whit- more, 1997; Holm and Richards, 2013). Moreover, the convergence of the New Britain and Trobriand Trenches presents an evident trend of eastward migration (Crook, 1989; Silver., 1991). In the eastern part of the Solomon Sea, the New Britain Trench parallels the distribution direction of the Solomon Islands in the southeast direction, whereas the Trobriand Trench merges with the Nubra fault, which is characterized by a right-lateral strike slip (Cameron, 2014). The Woodlark Basin on the south- ern side of the Nubara fault, which is characterized by an average water depth of approximately 3000m, is a young sea basin, and it began to spread 6Myr ago (Little., 2007). According to seismological studies (Yoneshima.,2005), the Woodlark Basin subducted beneath the Solomon Islands along the San Cristobal Trench, and the spreading center of the basin extended into the region at the northeast side of the Papuan Peninsula. The Woodlark Basin is a special area where the young sea basin exhibits a strong subduction activity with rapid expansion (Weissel., 1982). The Pocklington Trough is on the southern side of the Woodlark Basin and represents a relict trench that accommodated the north-dipping subduction of the Australian Plate beneath New Guinea (Holm., 2016). During the Oligocene, with the complete subduction of the Pocklington Basin (Schellart., 2006), the collision between New Guinea and the passive continental margin of the Australian Plate was initiated. The subduction along the Pocklington Trough gradually became inactive during the mid-late Miocene (12Myr) (Holm., 2015).

    Fig.1 Structural map showing the tectonic setting and geological tectonic units within the PN-SL. The top-right inset is a geographic map of the West Pacific and surrounding regions, in which a yellow rectangle shows the location of the study area, and the orange arrows imply the plate motion. The red lines denote the vertical sections of the tomography results in Fig.9. The gray dashed rectangle presents the location of Fig.2. The global positioning system (GPS) velocity indicates the Australia-Pacific Plate relative velocity from the United States Geological Survey. The bathymetry data were from the General Bathymetric Chart of the Oceans. The faults and relative GPS velocities of different blocks were mainly from the works of Cooper and Taylor (1985), Pegler et al. (1995), Wallace et al. (2004), Baldwin et al. (2012), and Holm et al. (2016). The black solid and dashed lines denote the currently active and inactive faults, respectively. MI, Malaita Island; NBI, New Britain Island; NBS, North Bismarck Sea; NBT, New Britain Trench; NGI, New Guinea Island; NGT, New Guinea Trench; NF, Nubara Fault; NST, North Solomon Trench; PP, Papuan Peninsula; PT, Pocklington Trough; SBS, South Bismarck Sea; SCT, San Cristobal Trench; SI, Solomon Islands; SS, Solomon Sea; TT, Trobriand Trench; WB, Woodlark Basin; MB, Manus Basin; WMT, West Melanesian Trench.

    The Ontong Java Plateau is located on the eastern side of the PN-SL, and the plateau summit is approximately 2000–2500m above the ocean floor. This plateau was created by two stages of magma activity from 122Myr to 90Myr ago and is the world’s largest oceanic plateau, with an area of 1.9×106km2(Mahoney., 1993; Taylor, 2006; Hanyu., 2017; Zhang., 2020). According to the seismic reflection data collected from the area of the Ontong Java Plateau and the area southwest across the North Solomon Trench, the average crustal thickness of the plateau can reach 30km (a maximum of 38km), which is 3–4 times thicker than the normal oceanic crust (Richardson., 2000; Miura., 2004; Inoue., 2008). The thick oceanic plateau is difficult to submerge in the trench, and the strong buoyancy induced by the thick crust-mantle of the plateau has a significant retarding effect on the plate subduction process, which can promote the collision between the plateau and the PN-SL.

    The Bismarck Sea, with an average water depth of near- ly 2000m, is the back-arc basin corresponding to the New Britain Trench. The record of submarine magnetic anomalies in the Bismarck Sea indicates a seafloor spreading that has been occurring since the Pliocene (Holm, 2016). The Bismarck Sea back-arc basin is composed of two microplates (North Bismarck and South Bismarck). The active center of the basin is now concentrated in the Manus Basin, whichhas expanded asymmetrically since 3.5Myr ago (Taylor, 1979; Both., 1986; Binns and Scott, 1993; Zhao., 2014; Ma., 2017). The West Melanesian Trench is on the north side of the Bismarck Sea and connected to the North Solomon Trench, which is the subduction area of the Pacific Plate during the early Cenozoic. The West Melanesian-North Solomon Trench is currently inactive or slightly active and closely related to the subduction of the Ontong Java Plateau (Schellart., 2006; Gong., 2019a).

    3 Data and Methods

    The P-wave traveltime data used in this study were obtained from the seismic catalog of the International Seismological Centre (ISC) from January 1960 to January 2017and corrected using the Engdahl-van der Hilst-Buland me- thod (Engdahl., 1998). The range of the inversion area is 138?–162?E and 0?–12?S (black rectangle in Fig.2). As published on the ISC website, this region has 82 seismic stations, and their distribution is shown in Fig.2. Given the absence of stations in the ocean, the spatial distribution of the seismic stations is limited by the distribution of islands in the study area.

    The teleseismic P-wave travel times used in this study were those from the seismic traveltime catalog published on the ISC website. The criteria for earthquake selection were as follows: 1) the epicentral distance from the center of the study area is between 30? and 90? to avoid in?uences on the travel time of the seismic rays in the core-mantle boundary and the upper mantle outside the study region (Zhao., 1994, 2013); 2) the magnitude of the earthquake is greater than 5.0; 3) the number of seismic stations in the study region recording each seismic event is greater than or equal to five. In accordance with the above criteria, 2011 teleseismic events and 15009 P-wave arrival times were selected. Fig.3 shows the distribution of these seismic events, which were distributed around the study region, which has a good back azimuth coverage.

    Fig.2 Distribution of seismic stations in the study region. The green dots represent the seismic station locations.

    Fig.3 Epicentral distribution of the teleseismic events used in this study from the ISC website. The colored dots represent the epicenters. The dot size denotes the event magnitude, whereas the dot color indicates the event depth. The event magnitude and depth follow the scales on the bottom and right sides, respectively, of the figure. The red dashed rectangle indicates the center of the study area.

    The one-dimensional (1-D) IASP91 velocity model was used to calculate the theoretical travel time of the teleseis- mic P-waves (Fig.4) and obtain the absolute traveltime re- siduals of each ray. In teleseismic tomography, relative tra- veltime residuals are generally used to avoid the influences of hypocenter mislocation, origin times, and complex structures outside the study region (Zhao., 1994), that is, the absolute traveltime residuals with the average traveltime residuals of all stations for each event in the area are removed. The traveltime residuals were corrected for crustal heterogeneity and topography. The crust model CRUST 1.0 model (http://igppweb.ucsd.edu/~gabi/crust1. html) was used in this process.

    The teleseismic tomography used in this study applied the shooting ray tracing method (Julian and Gubbins, 1977; Rawlinson., 2006; Zhang., 2014, 2018) to calculate the travel time and ray paths; this process is a traditional ray tracing method and is commonly used in seismic tomography. The 3-D ray tracing was conducted following Snell’s law, and by constantly adjusting and modifying the initial values of the ray parameters, relatively accurate ray paths and travel times can ultimately be obtained. The inversion problem of teleseismic tomography was solved by iteratively applying the damped least-squares algorithm (LSQR; Paige and Saunders, 1982), which can solve huge and sparse matrix equations with relatively less computation.

    Fig.4 1-D IASP91 velocity model.

    4 Reliability Analysis of Inversion Results

    The 1-D IASP91 (Kennett and Engdahl, 1991) velocity model was used as the initial velocity model (Fig.4). For the initial grid division, considering that the coverage density of the seismic stations in the area is relatively low, after numerous tests, we decided that the optimum grid interval in the longitudinal and latitudinal directions was 1? for our dataset. In the vertical direction, the grid interval was 100km, and the maximum depth was 900km.

    Given the multi-solution problem of geophysical inversion, the reliability of inversion needs to be verified and evaluated. To test the resolution of the observation system composed of the stations and seismic events used in this study, we conducted a checkerboard resolution test (CRT). In the CRT, the seismic stations, teleseismic events, and ray path distribution were completely consistent with the actual inversion. A 3-D perturbed velocity model was generated by adding the positive and negative velocity perturbations to an initial velocity model, and it used the stations and event distribution described above to calculate the theoretical relative travel time residuals of each ray for the inversion. By comparing the perturbed model and its inversion results, the scale of the smallest velocity anomaly that can be distinguished in the inversion results is the resolution of the model space.

    The CRT velocity model used in this study was based on the initial IASP91 model with added±3% perturbations and 1% random noise.repeated experiments on the scale of the velocity anomaly, when the scale of the abnormal block was 2?×2?×180km, the results of the CRT became ideal. Fig.5 shows the horizontal slices of the CRT results at different depths from 100km to 700km. Com- pared with the output and input models, the resolution in the shallow (<100km) study region was relatively low, and only in the middle part of the study region, near the New Britain Island and New Guinea Peninsula, was the velocity perturbation recovered due to the seismic station distribution, which severely limited the shallow ray path. With the increase in depth, most of the study region exhibited a better resolution, especially at depths greater than 300km.

    Fig.5 Results of the CRT. The recovered checkerboard results are shown in horizontal slices. The black dots indicate the seismic station locations.

    The seismic stations are primarily distributed on the islands in the middle western part of the study region (141?– 154?E), whereas the eastern part of the study region is mostly sea and therefore lacks seismic stations, which resulted in an improved ray coverage in the middle and western parts of the study region. Therefore, the resolutionof the middle and western parts (>154?E) of the study region was better than that of the eastern part at all depths.

    The inversion results of the LSQR algorithm were affected by the values of the damping parameter, which con- trols the convergence speed of the solution and smoothness of the model. By comparing the inversion results of different damping factor values, we obtained the relationship between the root mean square of the traveltime residuals and the smoothness of the model (Fig.6). When the damping factor was 20, the model smoothness and root mean square traveltime residuals were well balanced. There- fore, 20 was selected as the optimal value of the damping factor. Fig.7 shows the statistical results of the relative traveltime residuals before and after the inversion with a damping factor value of 20. Most of the residuals prior to the inversion were between ?1.5 and 1.5s. After the inversion, the residual distribution shrunk to the middle of the distribution, with the main body concentrated between ?1 and 1s. The distribution pattern conformed to the cha- racteristics of a normal distribution, and the root mean square of the traveltime residual decreased by approximately 20%, indicating that the velocity model after the inversion can match the observed relative traveltime residuals.

    Fig.6 Scheme used to estimate the optimum damping para- meter. The trade-off curve of the residual variance versus the model roughness shows that the optimal value ranges from 17 to 25. We selected the value of 20 as the optimal choice.

    Fig.7 Frequency histograms showing the magnitude and distribution of the relative traveltime residuals associated with the (a) initial and (b) ?nal models.

    5 Results

    With the 1-D IASP91 velocity model as the initial reference model, the velocity structure beneath the PN-SL was obtained using teleseismic traveltime tomography. Fig.8 shows the imaging results at various depth levels.

    In the depth range of 100–700km, the northwestern and northeastern sides of the Solomon Sea were characterized by high-velocity anomaly zones. The spatial distri- bution of the high-velocity anomaly zones presented a clear correspondence with the New Britain Trench. With the increase in depth, the high-velocity anomaly zone at the New Britain Trench gradually migrated northward. The high-velocity anomaly zone along the New Britain Trench became wider at a depth of 400km, and the southern boun- dary of the anomaly zone gradually expanded toward the Solomon Sea. At a depth of 700km, the linear-shaped high- velocity anomaly zone at the New Britain Trench was no longer complete, but the PN-SL presented a mixture of high- and low-velocity anomalies dominated by the high- velocity anomalies.

    Fig.8 Horizontal slices of tomography results. The velocity perturbation scale is shown at the bottom of the figure. The main geological structures represented by the black curves are the same as those in Fig.1.

    The Bismarck Sea is an evident low-velocity anomaly area, and this low-velocity anomaly can be traced to a depth of 600km. At the West Melanesian Trench north of the Bismarck Sea, remarkable high-velocity anomalies were distributed in the depth range of 100–200km; these were gradually dominated by low-velocity anomalies as the dep- th increased. At the Trobriand Trench, a high-velocity ano- maly zone can be observed in the velocity slices at 100–200km. Although the high-velocity anomaly was still ob- served near the Trobriand Trench as the depth increased, its spatial distribution pattern was not limited to the region along the Trobriand Trench. Evident low-velocity ano- malies were recorded within the area of the Woodlark Basin. In particular, the western part of the basin showed significant characteristics of low-velocity anomalies in the range of 400–600km. In the area of New Guinea Island, a remarkable high-velocity anomaly can be observed at a depth of 100km and can be traced down to a depth of 700km. Although this high-velocity anomaly within the depth range of 400–500km is divided by a low-velocity anomaly, the New Guinea Island still exhibits velocity cha- racteristics dominated by high-velocity anomalies.

    6 Discussion

    Based on the above analysis of seismic inversion results, we extracted four vertical sections along the red lines in Fig.1 to determine the deep shape of the subducted slab beneath the New Britain Trench and the eastern part of the New Guinea Island using the abundant distribution of the epicenters and good resolution (Fig.5). On the basis of the analysis of the deep-velocity anomalies beneath the PN-SL, we discussed the subduction pattern of the Solomon Sea Plate and its spatial geometric relationship with the subducted Pacific and Australian Plates.

    As shown in Fig.9, as the most active subduction system of the PN-SL, the velocity structure of the subducted Solomon Sea Plate presented significant spatial differences. At the New Britain Trench, the subduction depth of the Solomon Sea Plate increased gradually from west to east. The subduction depth was 300km on the P1 profile, which is similar to the findings of Peglerand Woodhouse (1995), and it increased significantly to 600km on the P2 and P3 profiles, which is also supported by the spatial distribution of the epicenters (Holm and Richards, 2013). The de- tached slabalso extended to a depth of 800km on the P4 profile. However, the shape of the detached slab was unclear under the constraints of the number of seismic events. This finding indicated that the width of the Solomon Sea Basin prior to subduction could have been at least 1000 km. If the Solomon Sea had started being consumed by the New Britain Trench 8Myr ago (Holm., 2016), the subduction rate of the slab would have been 10cmyr?1. As indicated by the high-velocity anomalies, the subduction angle of the Solomon Sea Plate shared a similar spatial variation trend with the subduction depth of the slab and gradually increased from west to east. The subduction angle of the slab increased from 45? on the P1 profile to 75? on the P2 and P3 profiles. On the P4 profile, the subducted slab was nearly vertical, which is similar to the subduction pattern of the old Pacific Plate at the Mariana Trench (., Jaxybulatov., 2013; Barklage., 2015; Wu., 2016; Gong., 2019b). Confusingly, the subduction pattern at the New Britain Trench was completely different from that at the Peru-Chile Trench (., Anderson., 2007; Hu., 2016), which is characterized by a similar seafloor age and subduction rate as the New Britain Trench (Müller., 2008; Schellart., 2011). In previous studies, the geometry of the subducting slab was interpreted to be controlled by its own gravitational pull (Peglerand Woodhouse, 1995), dia- chronous subduction initiation, and the eastern Solomon Sea remaining partly coupled to the Pacific (Hall and Spak- man, 2002). For the New Britain Trench, with its young subducted plate and rapid subduction rate, we proposed that the high angle of the slab may be attributed to the migration of the trench and mantle convection (., Faccenna., 2007; Boutelier and Cruden, 2008).

    Fig.9 Vertical sections of tomography results along the profiles shown in Fig.1. The velocity perturbation scale is shown at the bottom of the figure. L1, L2, and L3 denote three primary low-velocity anomalies. The gray dashed lines represent the 410 and 660km-depth lines. The upper image shows the location of each cross-section. WMT, West Melanesian Trench; NST, North Solomon Trench; TT, Trobriand Trench.

    In contrast to the subduction at the New Britain Trench, the subduction depth of the Solomon Sea Plate at the Tro- briand Trench changed from 200km along the P1 profile to 100km along the P4 profile. The eastward weakening of the subduction activity was supported by spatial changes in the seismic intensity, with a decrease in the seismic events from the P1 to P4 profiles. In addition, the subduction angle of the Solomon Sea Plate at the Trobriand Trench was 45?. Overall, the subduction activity of the Solomon Sea Plate at the Trobriand Trench was weaker than that at the New Britain Trench. According to Lock. (1987), the multichannel seismic reflection data showed a gentle onlap of the uppermost trench sediment onto the frontal anticline at the toe of the accretionary wedge. Moreover, the small earthquakes, shallow subduction depth, and lack of arc magmatism suggest that the subduction zone at the Trobriand Trench is presently inactive and does not accommodate a major component of the regional convergence (Davies., 1984; Wallace., 2004; Holm., 2016). Along with the collision between the Ontong Java Plateau, the subduction of the Solomon Sea Basin along the New Britain and Trobriand Trenches is the main adjustment style to accommodating the convergence of the plates. Different from the active and strong subduction along the New Britain Trench, the subduction activity at the Trobriand Trench weakened at the early Pliocene and may be stagnant at present. The different times of subduction activity led to the heterogeneous slab patterns in the two trenches. Further, as shown on the P1 profile, the Solomon Sea Plate formed an ‘inverted U-shaped’ double-dipping structural pattern at the Trobriand Trench and New Britain Trench (Pegler., 1995). As indicated by the focal mechanism and stress field analysis (Woodhead., 2010; Holm and Richards, 2013), the double subduction of the Solomon Sea Plate may be completely decoupled from the overlying Australian Plate. Thus, the Solomon Sea Plate at this location has been completely consumed beneath the convergence zone.

    As shown in Fig.9, the deep structure of the early subduction system and its geometric relationship with the present active subducted Solomon Sea Plate is slightly more complicated. The Pacific Plate subducted strongly in the Cretaceous period along the West Melanesian and North Solomon Trenches, and the subduction of the Pacific Plate became extremely weak following the collision between the Ontong Java Plateau and Solomon Islands during the Oligocene period (Petterson., 1997). According to the inverted velocity model, the early subduction structure of the Pacific Plate showed spatial differences under the influence of later subduction activities. On the P1 profile, the Pacific slab extended to a depth of 200km, and no slab was present in the deep mantle. According to the P2, P3, and P4 profiles, the late subduction of the Solomon Sea Plate had been intercalated within the early subduction of the Pacific Plate. Given the influence of pre- existing faults within the slab and the late subduction activity, the Pacific slab in the deep mantle is discontinuous. However, the Pacific slab subducted downward at a large angle of >60? and is stagnant in the mantle transition zone. The P2 profile showed that the detached Pacific Plate extended to a depth of 800km. This finding is consistent with that of the inverted velocity model from the work of Hall and Spakman (2002), and our inversion results showed a considerably clearer shape of the subducted slab. As indicated by the P2, P3, and P4 profiles, at least 1000km of the Pacific Plate has been consumed in the deep mantle at the West Melanesian Trench.

    The Pocklington Trough in the southern part of the PN- SL was active during the mid-late Eocene (45Myr ago) and was inactive from 12Myr ago (Schellart., 2006; Holm., 2015). On the P1 profile, we observed that the Australian Plate has a similar subduction scale to the eastern Pacific Plate. The Australian Plate extended to a depth of 600km at an average dip of 60?, and 800km of the slab has been consumed in the mantle. Although no similar interpenetrating relationship exists between the Pacific Plate and Solomon Sea Plate, the subducted Australian Plate collided with the subducted Solomon Sea Plate at the Trobriand Trench at a depth of 200km. Further, the Australian slab is characterized by a break-off within the depth range of 300–400km, which is evidenced by outcropped igneous rocks (Holm., 2015). Based on the tomographic inversion result and petrological records, we proposed that the Australian slab break-off may be correlated to a subduction polarity reversal caused by the south- ward subduction of the Solomon Sea Plate at the Trobriand Trench. This reversal would have removed the mantle wedge beneath the arc and culminated in ophiolite obduction in the Papuan Peninsula (Fitz and Mann, 2013).

    In addition, the velocity profiles showed evident low- velocity anomalies. According to our comprehensive analy- sis, three primary low-velocity anomalies, namely L1, L2, and L3, were observed (Fig.9). The L1 low-velocity anoma- ly was located between the subducted Solomon Sea Plate and the Pacific Plate and primarily distributed within the mantle above the Solomon Sea Plate. This anomaly was likely caused by the dehydration of the subducted plate, which led to the melting of the mantle, and should represent the deep dynamic factor of the Pliocene expansion of the Bismarck Sea back-arc basin with respect to the New Britain Trench (Baldwin., 2012; Holm., 2016).

    The L2 low-velocity anomalywas located beneath the Solomon Sea and above the subducted Pacific Plate and primarily distributed at depths of 200–400km. In terms of the current spatial distribution, the anomaly may be primarily caused by the melting of mantle induced by the dehydration of the subducted Pacific Plate, which in turn led to the expansion of the Solomon Sea. Thus, the expansion of the Solomon Sea is a ‘back-arc expansion basin of a single subduction system’ induced by the Pacific Plate subduction (Honza., 1987; Hall and Sparkman, 2003) rather than a back-arc basin of a double subduction system (Schellart., 2006). In addition, the expansion of the Solomon Sea may not be closely related to the sub- duction activity of the Australian Plate along the Pocklington Trough. Notably, a key issue was neglected in the above analysis: the early subducted Australian Plate might have been modified by later tectonic activities. In parti- cular, the spread of the Woodlark Basin since the Pliocene induced the migration of the western Pocklington Trough and the deep subduction of the Australian Plate, which can be observed in the velocity structure shown on the P1 profile. As indicated by the P1 profile, which is on the western edge of the Solomon Sea and does not cross the Solomon Sea, the subducted Australian Plate lies beneath the double-subducted Solomon Sea Plate. In addition, a significant low-velocity anomaly body exists between the subducted Australian Plate and the double-subducted Solomon Sea Plate. From the tectonic location, this structure may be the same as that in the L2 low-velocity anomaly in the east, which dominated the back-arc expansion of the Solomon Sea. The subduction of the Australian Plate also played a role in the expansion of the Solomon Sea Ba- sin. As for the coupling mode of the Solomon Sea spreading with the subduction of the Pacific and Australian Plate, further research is needed, in particular on the subduction dynamic simulations in the PN-SL.

    The L3 low-velocity anomaly was located in the deep zone south of the Trobriand Trench and shown on the P4 profile. From the P4 profile, the L3 low-velocity anomaly extended from at least 100 km down to a depth of 600km, with a scale larger than that of the L2 low-velocity anoma- ly. Using the seismic data recorded by the fixed stations, Abers. (2002) observed a significant low-velocity anomaly within the depth range of 0–240km south of the Trobriand Trench; its position is consistent with the L3 low-velocity anomaly inverted in this paper. To a certain extent, the study is a supplement to the work of Abers. (2002), revealing the deep distribution of the L3 low-velocity anomaly. Combined with the study of normal focal mechanisms, young magnetic anomalies, and rapid exhumation (>1cmyr?1) of Pliocene eclogite at the D’Entrecasteaux Islands (Weissel, 1982; Crook and Taylor, 1994; Taylor., 1999; Baldwin., 2004; Little., 2007), the northward extension of the Woodlark Basin is closely related to the generation of the large- scale L3 low-velocity anomaly. The generation of the L3 low-velocity anomaly provided a dynamic factor for the expansion of the Woodlark Basin. As to whether the asthenospheric mantle corresponding to the L3 low-velocity anomaly body is an active or passive upwelling, or whe- ther the Woodlark Basin was caused by back-arc expansion (Fitz and Mann, 2013), the activities of a tectonic block beneath the Goodenough Bay (Wallace., 2014), or the subduction and rotation of the Solomon Sea Plate (Benes., 1994; Baldwin., 2012; Webb., 2014), this study alone may be insufficient to provide a definite answer. In addition, the L3 low-velocity anomaly was found on the southern side of the subducted Pacific Plate with respect to the P4 profile. Thus, the L3 low- velocity anomaly may not be related to the subduction of the Pacific Plate. Combining the horizontal slices of the tomography results and the P4 profile (Figs.8 and 9), the formation of the L3 low-velocity anomaly may be dominated by the subduction of the Australian Plate at the Pock- lington Trough.

    7 Conclusions

    Through the teleseismic P-wave tomography of the PN-SL, multiple and diverse types of subduction systems induced the complex deep structure beneath the PN-SL. As the most active subduction system of the PN-SL, the young Solomon Sea Plate is characterized by typical and distinct spatial differences in its subduction patterns, which cannot be revealed by previous studies. As the subduction angle of the Solomon Sea Plate increased from west to east along the New Britain Trench, the subduction depth also deepened, extending to a depth of 800km. At the Trobriand Trench, the subduction depth of the Solomon Sea Plate is less than 200km, and the subduction scale is significantly smaller than that of the New Britain Trench on the northern side of the Solomon Sea. Under the background of strong convergence of the Australian and Pacific Plates, the subduction rate of the Solomon Sea slab is close to or more than 10cmyr?1since the late Miocene. During its large-scale subduction, the Solomon Sea Plate strongly changed the subduction structures of the early subducted Pacific and Australian Plates, and the subduction plates intersected each other. In addition, the subduction of the Solomon Sea Plate induced mantle melting and led to the formation of low-velocity anomaly bodies, which directly became a deep dynamic source for the expansion of the back-arc sea basin. As for the genetic mechanism of the expansion of the Solomon Sea back-arc basin, the inverted velocity structure indicates that the early subducted Pacific Plate played a key role in its expansion. However, constrained by the influence of the late tectonic transformation and general resolution of the velocity structure within the deep mantle, an important role of the Australian Plate in the expansion of the Solomon Sea cannot be completely excluded. Thus, the double subduction systems possibly dominated the formation of the Solomon Sea back-arc basin. Further, the subduction of the Australian Plate may have induced the formation of the low- velocity anomaly below the Woodlark Basin. As for whe- ther the expansion of the Woodlark Basin was caused by the upwelling of mantle melt or the movement of the micro- plate, further studies in combination with regional structural numerical simulations and other aspects are necessary.

    Acknowledgements

    We thank Wessel & Smith (https://www.soest.hawaii.edu/ gmt/) for the free use of GMT software, which was used to produce most of the ?gures used in the study. We are grateful for the algorithmic support of the traveltime tomography study from Dr. Fengxue Zhang at the Instituteof Geophysics, China Earthquake Administration. This work was supported by the National Natural Science Foundation of China (Nos. 91858215 and 41906048), the Fundamental Research Funds for the Central Universities (No. 201964015), and Laboratory for Marine Mineral Resources, Qingdao National Laboratory for Marine Science and Technology (No. MMRZZ201801).

    Abers, G. A., Ferris, A., Craig, M., Davies, H., Lerner-Lam, A. L., Mutter, J. C.,., 2002. Mantle compensation of active metamorphic core complexes at Woodlark rift in Papua New Guinea., 418: 862-865.

    Abers, G. A., and Roecker, S. W., 1991. Deep structure of an arc- continent collision: Earthquake relocation and inversion for upper mantle P and S wave velocities beneath Papua NewGuinea., 96 (B4): 6379-6401.

    Anderson, M., Alvarado, P., Zandt, G., and Beck, S., 2007. Geo- metry and brittle deformation of the subducting Nazca Plate, Central Chile and Argentina., 171 (1): 419-434.

    Baldwin, S. L., Monteleone, B. D., Webb, L. E., Fitzgerald, P. G., Grove, M., and Hill, E. J., 2004. Pliocene eclogite exhumation at plate tectonic rates in eastern Papua New Guinea., 431: 263-267.

    Baldwin, S. L., Fitzgerald, P. G., and Webb, L. E., 2012. Tecto- nics of the New Guinea Region., 40: 495-520.

    Barklage, M., Wiens, D. A., Conder, J. A., Pozgay, S., Shiobara, H., and Sugioka, H., 2015. P and S velocity tomography of the Mariana subduction system from a combined land-sea seis- mic deployment., 16 (3): 681-704.

    Benes, V., Scott, S. D., and Binns, R. A., 1994. Tectonics of rift propagation into a continental margin: Western Woodlark Basin, Papua New Guinea., 99 (B3): 4439-4455.

    Binns, R. A., and Scott, S. D., 1993. Activity forming polymer- tallic sulfide deposits associated with felsic volcanic rocks in the eastern Manus back-arc basin, Papua New Guinea., 88 (8): 2226-2236.

    Both, R., Crook, K., Taylor, B., Brogan, S., Chappell, B., Frankel, E.,., 1986. Hydrothermal chimneys and associated fauna in the Manus back-arc basin, Papua New Guinea., 67 (21): 489-490.

    Boutelier, D. A., and Cruden, A. R., 2008. Impact of regional mantle ?ow on subducting plate geometry and interplate stress: Insights from physical modelling., 174 (2): 719-732.

    Cameron, M. L., 2014. Rifting and subduction in the Papuan Peninsula, Papua New Guinea: The significance of the Trobriand tough, the Nubara strike-slip fault, and the Woodlark rift to the present configuration of Papua New Guinea. PhD thesis. The University of Alabama.

    Cooper, P. A., and Taylor, B., 1985. Polarity reversal in the Solo- mon Islands arc., 314: 428-430.

    Cooper, P., and Taylor, B., 1987. Seismotectonics of New Guinea: A model for arc reversal following arc-continent collision., 6 (1): 53-67.

    Crawford, A. J., Meffre, S., and Symonds, P. A., 2003. 120 to 0 Ma tectonic evolution of the Southwest Pacific and analogous geological evolution of the 600 to 220Ma Tasman Fold Belt System., 22: 377-397.

    Crook, K. A. W., 1989. Suturing history of an allochthonous terrane at a modern plate boundary traced by flysch-to-molasse facies transitions., 61 (1-2): 49-79.

    Crook, K. A. W., and Taylor, B., 1994. Structure and Quaternary tectonic history of the Woodlark triple junction region, Solomon Islands., 16 (1): 65-89.

    Davies, H. L., Symonds, P. A., and Ripper, I. D., 1984. Structure and evolution of the southern Solomon Sea region., 9: 49-68.

    DeMets, C., Gordon, R. G., and Argus, D. F., 2010. Geologically current plate motions., 181 (1): 1-80.

    Engdahl, E. R., van der Hilst, R. D., and Buland, R., 1998. Global teleseismic earthquake relocation with improved travel times and procedures for depth determination., 88 (3): 722-743.

    Faccenna, C., Heuret, A., Funiciello, F., Lallemand, S., and Becker, T. W., 2007. Predicting trench and plate motion from the dynamics of a strong slab., 257 (1-2): 29-36.

    Finlayson, D. M., Cull, J. P., Wiebenga, W. A., Furumoto, A. S., and Webb, J. P., 1972. New Britain-New Ireland crustal seismic refraction investigations 1967 and 1969., 29 (3): 245-253.

    Fitz, G., and Mann, P., 2013. Tectonic uplift mechanism of the Goodenough and Fergusson Island gneiss domes, eastern Pa- pua New Guinea: Constraints from seismic reflection and well data., 14 (10): 3969- 3995.

    Gong, W., Jiang, X., Xing, J., Li, D., and Xu, C., 2019a. Subduction dynamics of the New-Guinea-Solomon arc system: Constraints from the subduction initiation of the plate., 39 (9): 115-129 (in Chinese with English abstract).

    Gong, W., Jiang, X., Xing, J., Xu, C., and Xu, X., 2019b. He- terogeneous strain regime at the west of the Ogasawara Plateau in the western Paci?c Ocean from inversion of earthquake focal mechanisms., 180: 103868, DOI: 10.1016/j.jseaes.2019.103868.

    Hall, R., and Spakman, W., 2002. Subducted slabs beneath the eastern Indonesia-Tonga region: Insights from tomography., 201 (2): 321-336.

    Hall, R., and Spakman, W., 2003. Mantle structure and tectonic evolution of the region north and east of Australia 22., 372: 361-381.

    Hanyu, T., Tejada, M. L. G., Shimizu, K., Ishizuka, O., Fujii, T., Kimura, J.,., 2017. Collision-induced post-plateau volcanism: Evidence from a seamount on Ontong Java Plateau., 294-295: 87-96.

    Holm, R. J., and Richards, S. W., 2013. A re-evaluation of arc- continent collision and along-arc variation in the Bismarck Sea region, Papua New Guinea., 60 (5): 605-619.

    Holm, R. J., Rosenbaum, G., and Richards, S. W., 2016. Post 8 Ma reconstruction of Papua New Guinea and Solomon Islands: Microplate tectonics in a convergent plate boundary setting., 156: 66-81.

    Holm, R. J., Spandler, C., and Richards, S. W., 2015. Continental collision, orogenesis and arc magmatism of the Miocene Maramuni arc, Papua New Guinea., 28 (3): 1117-1136.

    Honza, E., Davies, H. L., Keene, J. B., and Tiffin, D. L., 1987. Plate boundaries and evolution of the Solomon Sea region., 7 (3): 161-168.

    Honza, E., Miyazaki, T., and Lock, J., 1989. Subduction erosion and accretion in the Solomon Sea region., 160 (1-4): 49-62.

    Hu, J., Liu, L., Hermosillo, A., and Zhou, Q., 2016. Simulation of late Cenozoic South American flat-slab subduction using geodynamic models with data assimilation., 438: 1-13.

    Inoue, H., Coffin, M. F., Nakamura, Y., Mochizuki, K., andKroenke, L. W., 2008. Intrabasement reflections of the Ontong Java Plateau: Implications for plateau construction., 9 (4): Q04014, DOI: 10.1029/ 2007GC001780.

    Jaxybulatov, K., Koulakov, I., and Dobretsov, N. L., 2013. Seg- mentation of the Izu-Bonin and Mariana slabs based on the analysis of the Benioff seismicity distribution and regional tomography results., 4 (1): 59-73.

    Joshima, M., Okuda, Y., Murakami, F., Kishimoto, K., and Honza, E., 1987. Age of the Solomon Sea Basin from magnetic lineations., 6: 229-234.

    Julian, B. R., and Gubbins, D., 1977. Three-dimensional seismic ray tracing., 43 (1): 95-113.

    Kennett, B. L. N., and Engdahl, E. R., 1991. Traveltimes for glo- bal earthquake location and phase identi?cation., 105 (2): 429-465.

    Li, S., Suo, Y., Liu, B., Liu, Y., Li, X., Zhao, S.,., 2018. Microplate tectonics theory: Insights from microblocks in the global oceans and continental margins., 25 (5): 324-355 (in Chinese with English abstract).

    Little, T. A., Baldwin, S. L., Fitzgerald, P. G., and Monteleone, B., 2007. Continental rifting and metamorphic core complex formation ahead of the Woodlark spreading ridge, D’Entre- casteaux Islands, Papua New Guinea., 26 (1): TC 1002, DOI: 10.1029/2005TC001911.

    Lock, J., Davies, H. L., Tiffin, D. L., Murakami, F., and Kisimoto, K., 1987. The Trobriand subduction system in the western Solomon Sea., 7: 129-134.

    Ma, Y., Zeng, Z. G., Chen, S., Yin, X., and Wang, X., 2017. Origin of the volcanic rocks erupted in the eastern Manus Basin: Basaltic andesite-andesite-dacite associations., 16 (3): 389-402.

    Mahoney, J. J., Storey, M., Duncan, R. A., Spencer, K. J., and Pringle, M., 1993. Geochemistry and age of the Ontong Java Plateau. In:. Pringle, M. S.,., eds., Wiley, Washington D. C., 233-262.

    Mann, P., and Taira, A., 2004. Global tectonic significance of the Solomon Islands and Ontong Java Plateau convergent zone., 389 (3-4): 137-190.

    Miura, S., Suyehiro, K., Shinohara, M., Takahashi, N., Araki, E., and Taira, A., 2004. Seismological structure and implications of collision between the Ontong Java Plateau and Solomon Island arc from ocean bottom seismometer-airgun data., 389 (3-4): 191-220.

    Müller, R. D., Sdrolias, M., Gaina, C., and Roest, W. R., 2008. Age, spreading rates, and spreading asymmetry of the world’s ocean crust., 9 (4): Q0 4006.

    Paige, C. C., and Saunders, M. A., 1982. LSQR: An algorithm for sparse linear equations and sparse least squares., 8: 43-71.

    Pegler, G., Das, S., and Woodhouse, J. H., 1995. A seismological study of the eastern New Guinea and the western Solomon Sea regions and its tectonic implications., 122 (3): 961-981.

    Petterson, M. G., Neal, C. R., Mahoney, J. J., Kroenke, L. W., Saunders, A. D., Babbs, T. L.,., 1997. Structure and deformation of north and central Malaita, Solomon Islands: Tectonic implications for the Ontong Java Plateau-Solomon arc collision, and for the fate of oceanic plateaus., 283 (1-4): 1-33.

    Rawlinson, N., Reading, A. M., and Kennett, B. L. N., 2006. Lithospheric structure of Tasmania from a novel form of teleseismic tomography., 111 (B2): B02301, DOI: 10.1029/2005JB003803.

    Richardson, W. P., Okal, E. A., and Van der Lee, S., 2000. Ray- leigh-wave tomography of the Ontong-Java Plateau., 118 (1-2): 29-51.

    Schellart, W. P., Lister, G. S., and Toy, V. G., 2006. A late Cretaceous and Cenozoic reconstruction of the Southwest Pacific region: Tectonics controlled by subduction and slab rollback processes., 76 (3-4): 191-233.

    Schellart, W. P., and Spakman, W., 2015. Australian plate motion and topography linked to fossil New Guinea slab below Lake Eyre., 421: 107- 116.

    Schellart, W. P., Stegman, D. R., Farrington, R. J., and Moresi, L., 2011. Influence of lateral slab edge distance on plate velocity, trench velocity, and subduction partitioning., 116 (B10): B10408, DOI: 10.1029/20 11JB008535.

    Silver, E. A., Abbott, L. D., Kirchoff-Stein, K. S., Reed, D. L., and Bernstein-Taylor, B., 1991. Collision propagation in PapuaNew Guinea and the Solomon Sea., 10 (5): 863-874.

    Taylor, B., 1979. Bismarck Sea–Evolution of a back-arc basin., 7 (4): 171-174.

    Taylor, B., 2006. The single largest oceanic plateau: Ontong Java-Manihiki-Hikurangi., 241(3-4): 372-380.

    Taylor, B., Goodliffe, A. M., and Martinez, F., 1999. How continents break up: Insights from Papua New Guinea., 104 (B4): 7497-7512.

    Taylor, B., Goodliffe, A., Martinez, F., and Hey, R., 1995. Continental rifting and initial sea-floor spreading in the Woodlark Basin., 374: 534-537.

    Tregoning, P., Lambeck, K., Stolz, A., Morgan, P., McClusky, S. C., van ber Beek, P.,., 1998. Estimation of current plate motions in Papua New Guinea from global positioning system observation., 103 (B6): 12181-12203.

    Wallace, L. M., Ellis, S., Little, T., Tregoning, P., Palmer, N., Rosa, R.,., 2014. Continental breakup and UHP rock ex- humation in action: GPS results from the Woodlark Rift, Papua New Guinea., 15 (11): 4267-4290.

    Wallace, L. M., Stevens, C., Silver, E., McCaffrey, R., Loratung, W., Hasiata, S.,., 2004. GPS and seismological constraints on active tectonics and arc-continent collision in Papua New Guinea: Implications for mechanics of microplate rotations in a plate boundary zone., 109 (B5): B05404.

    Webb, L., Baldwin, S. L., and Fitzgerald, P. G., 2014. The early- middle Miocene subduction complex of the Louisiade Archipelago, southern margin of the Woodlark Rift., 15 (10): 4024-4046.

    Weissel, J. K., Taylor, B., and Karner, G. D., 1982. The opening of the Woodlark Basin, subducting of the Woodlark spreading system, and the evolution of northern Melanesia since mid- Pliocene time., 87 (1-4): 253-277.

    Whitmore, G. P., Johnson, D. P., Crook, K. A. W., Galewsky, J., and Silver, E. A., 1997. Convergent margin extension associated with arc-continent collision: The Finsch Deep, Papua New Guinea., 16 (1): 77-87.

    Woodhead, J., Hergt, J., Sandiford, M., and Johnson, W., 2010. The big crunch: Physical and chemical expressions of arc/ continent collision in the western Bismarck arc., 190 (1-2): 11-24.

    Wu, J., Suppe, J., Lu, R., and Kanda, R., 2016. Philippine Sea and East Asian Plate tectonics since 52Ma constrained by new subducted slab reconstruction methods., 121 (6): 4670-4741.

    Yoneshima, S., Mochizuki, K., Araki, E., Hino, R., Shinohara, M., and Suyehiro, K., 2005. Subduction of the Woodlark Basin at New Britain Trench, Solomon Islands region., 397 (3-4): 225-239.

    Zelt, B. C., Taylor, B., and Goodliffe, A. M., 2001. 3-D crustal velocity structure at the rift tip in the western Woodlark Basin., 28 (15): 3015-3018.

    Zhang, F. X., Wu, Q. J., and Ding, Z. F., 2018. A P-wave velo- city study beneath the eastern region of Tibetan Plateau and its implication for plateau growth., 63 (19): 1949-1961 (in Chinese with English abstract).

    Zhang, F. X., Wu, Q. J., and Li, Y. H., 2014. A traveltime tomography study by teleseismic S wave data in the Northeast China area., 57 (1): 88-101 (in Chinese with English abstract).

    Zhang, J., Luo, Y., and Chen, J., 2020. Oceanic Plateau formation implied by Ontong Java Plateau, Kerguelen Plateau and Shatsky Rise., 19 (2): 351-360.

    Zhao, D., Hasegawa, A., and Kanamori, H., 1994. Deep structure of Japan subduction zone as derived from local, regional, and teleseismic events., 99 (B11): 22313-22329.

    Zhao, D., Yamamoto, Y., and Yanada, T., 2013. Global mantle heterogeneity and its in?uence on teleseismic regional tomography., 23 (2): 595-616.

    Zhao, H., Zeng, Z., Yin, X., and Chen, S., 2014. Silicon and oxygen isotopic composition of igneous rocks from the eastern Manus Basin., 13 (3): 421-427.

    J. Ocean Univ. China(Oceanic and Coastal Sea Research)

    https://doi.org/10.1007/s11802-022-4626-x ISSN 1672-5182, 2022 21 (3): 694-706

    (June 10, 2020;

    February 21, 2021;

    March 3, 2022) ? Ocean University of China, Science Press and Springer-Verlag GmbH Germany 2022

    #The two authors contributed equally to this work.

    Corresponding author. E-mail: 21120411040@ouc.edu.cn

    (Edited by Chen Wenwen)

    永修县| 金湖县| 铁岭县| 汪清县| 万山特区| 阿坝| 普定县| 长宁区| 荣昌县| 松原市| 巴林左旗| 华亭县| 兴隆县| 平和县| 梓潼县| 东方市| 饶平县| 陈巴尔虎旗| 湘乡市| 新密市| 鸡东县| 开鲁县| 明星| 边坝县| 宁河县| 固阳县| 肥东县| 弋阳县| 徐汇区| 修水县| 宁城县| 玉屏| 康马县| 荣昌县| 龙川县| 原阳县| 龙泉市| 隆尧县| 麻阳| 固镇县| 沾化县|