Davood Shariatmadari, Seyed Mostafa Siadatmousavi, Cyrus Ershadi. Numerical study of power production from tidal energy in the Khuran Channel and its feedback on background hydrodynamics[J]. Acta Oceanologica Sinica, 2022, 41(5): 173-182. doi: 10.1007/s13131-021-1968-y
Citation: Weina Zhao, Zhiqiang Wu, Fanghui Hou, Xunhua Zhang, Tianyao Hao, Hanjoon Kim, Yanpeng Zheng, Shanshan Chen, Huigang Wang. Velocity structure in the South Yellow Sea basin based on first-arrival tomography of wide-angle seismic data and its geological implications[J]. Acta Oceanologica Sinica, 2023, 42(2): 104-119. doi: 10.1007/s13131-022-2028-y

Velocity structure in the South Yellow Sea basin based on first-arrival tomography of wide-angle seismic data and its geological implications

doi: 10.1007/s13131-022-2028-y
Funds:  The National Natural Science Foundation of China under contract No. 41806048; the Open Fund of the Hubei Key Laboratory of Marine Geological Resources under contract No. MGR202009; the Fund from the Key Laboratory of Deep-Earth Dynamics of Ministry of Natural Resource, Institute of Geology, Chinese Academy of Geological Sciences under contract No. J1901-16; the Aoshan Science and Technology Innovation Project of Pilot National Laboratory for Marine Science and Technology (Qingdao) under contract No. 2015ASKJ03-Seabed Resources; the Fund from the Korea Institute of Ocean Science and Technology (KIOST) under contract No. PE99741.
More Information
  • The South Yellow Sea basin is filled with Mesozoic–Cenozoic continental sediments overlying pre-Palaeozoic and Mesozoic–Palaeozoic marine sediments. Conventional multi-channel seismic data cannot describe the velocity structure of the marine residual basin in detail, leading to the lack of a deeper understanding of the distribution and lithology owing to strong energy shielding on the top interface of marine sediments. In this study, we present seismic tomography data from ocean bottom seismographs that describe the NEE-trending velocity distributions of the basin. The results indicate that strong velocity variations occur at shallow crustal levels. Horizontal velocity bodies show good correlation with surface geological features, and multi-layer features exist in the vertical velocity framework (depth: 0–10 km). The analyses of the velocity model, gravity data, magnetic data, multi-channel seismic profiles, and drilling data showed that high-velocity anomalies (>6.5 km/s) of small (thickness: 1–2 km) and large (thickness: >5 km) scales were caused by igneous complexes in the multi-layer structure, which were active during the Palaeogene. Possible locations of good Mesozoic and Palaeozoic marine strata are limited to the Central Uplift and the western part of the Northern Depression along the wide-angle ocean bottom seismograph array. Following the Indosinian movement, a strong compression existed in the Northern Depression during the extensional phase that caused the formation of folds in the middle of the survey line. This study is useful for reconstructing the regional tectonic evolution and delineating the distribution of the marine residual basin in the South Yellow Sea basin.
  • Oceanic fronts, or transition zones between water masses, play important roles in regulating oceanic heat, energy, and matter balances through associated vertical transport (Lévy et al., 2001; Ruiz et al., 2019). They also influence the atmospheric boundary layer (Xie, 2004). The observation and modeling of oceanic fronts are topics of interest in oceanography.

    The South China Sea (SCS) is a large marginal sea in the tropical western Pacific Ocean. The complex topography and monsoon allow multiscale oceanic structures to prevail in the SCS; these include branches of the Kuroshio Current and the SCS Warm Current as well as western boundary currents, mesoscale eddies (MEs), river plumes, upwelling, and submesoscale structures (Hu et al., 2012; Yuan et al., 2006; Zhong et al., 2017; Feng et al., 2020). Fronts have been detected in the above structures, i.e., Kuroshio fronts (Liu et al., 2017), upwelling fronts (Jing et al., 2015), and river plume fronts (Qiu et al., 2017a). Seasonal variations of above thermal fronts were first statistically revealed by Wang et al. (2001), who found that thermal fronts were strong in winter to spring and weak in summer to autumn. In addition to the above fronts, which have lifetimes longer than one month, there are other fronts with temporal scales less than 30 d in the SCS (Hosoda et al., 2012), which might be induced by MEs, filaments and so on.

    MEs are common in the SCS. The horizontal lengths/time scales of MEs are 50–300 km/1–10 months (Capet et al., 2008; McWilliams, 2016). Their horizontal scales of motion are characterized by baroclinic instability, at which the Rossby number ${Ro}\!=\!\dfrac{V}{fL}$<1, where $ V $ and $ L $ are horizontal velocity and length scales, respectively, and $\, f $ is the Coriolis frequency (Torres et al., 2018). Qiu et al. (2017a) found that the length scales depended on the local energy of MEs. One high eddy kinetic energy (EKE) band occurred in the northern SCS (NSCS), where both the number of MEs and the EKEs have no significant seasonal variations (Chen et al., 2009; Cheng and Qi, 2010). Interannual variations in EKE were suggested to be induced by the El Niño-Southern Oscillation (ENSO), which show a negative correlation between the SCS EKE and the ENSO index through conveying Kuroshio transport (Sun et al., 2016) or wind stress curl (Wang et al., 2008; Cheng and Qi, 2010). MEs deform with asymmetric shapes during propagation (Wang et al., 2018; Qiu et al., 2019b), resulting in strong vertical currents and turbulence at the ME boundary associated with thermal fronts (Yang et al., 2017, 2019; Qiu et al., 2019a).

    The signals of strong coastal and Kuroshio fronts in winter and strong upwelling and river plume fronts in summer hide the high-frequency frontal signals in seasonal variation studies (Wang et al., 2001; Jing et al., 2015; Qiu et al., 2017a). The SST fronts at the mesoscale eddy edge (hereafter, ME fronts) are expected to have high-frequency variations, because MEs easily deform and propagate at a mean speed of ~0.1 m/s (Chen et al., 2009; Su et al., 2020). As anticyclones have negative vorticity and cyclones have positive vorticity, the numbers of generated anticyclones and cyclones exhibit different seasonal variations due to monsoon-driven wind stress curl (Chen et al., 2009; Wang et al., 2008), which may influence the ME front. In this study, we examine the characteristics and possible mechanisms of ME fronts to better understand the interaction between MEs and SST fronts and improve our understanding of oceanic physical processes in the NSCS.

    The data and methods are presented in Section 2, ME fronts identified from the observation data are presented in Section 3, statistical analyses are presented in Section 4, possible mechanisms for ME fronts are presented in Section 5, and summaries are presented in Section 6.

    To investigate the structures at ME fronts, we collected observations from an underwater glider. The Chinese underwater glider was designed by the Shenyang Institute of Automation, Chinese Academy of Sciences, and named “Sea wing”. “Sea wing” underwater gliders have been successfully used to investigate sea surface cooling (Qiu et al., 2015), mesoscale eddies (Shu et al., 2016; Qiu et al., 2019b), and submesoscale eddies (Qiu et al., 2019a). The underwater glider captured 205 vertical profiles of temperature, salinity and pressure from April 19 to June 15, 2015. We interpolated the temperature and salinity to a 1-m vertical resolution.

    We used the HYbrid Coordinate Ocean Model (HYCOM), a data-assimilative hybrid isopycnal-sigma-pressure (generalized) coordinate ocean model, to calculate vertical velocity within the MEs. Descriptions of the data set are presented at https://www.hycom.org/. The data comprise 33 vertical levels at (1/12)° horizontal resolution. HYCOM data have been successfully used to observe MEs in the NSCS (Zhang et al., 2013; Huang et al., 2017).

    We used the operational SST and sea ice analysis (OSTIA) SST products to detect SST fronts. OSTIA SST products are a combination of microwaves, infrared satellite measurements, and in situ SST data available over global telecommunications system products (Donlon et al., 2011). Products from 2007 to 2017 with daily temporal resolution and ~5 km spatial resolution were downloaded from ftp://ftp.nodc.noaa.gov/pub/data.nodc/ghrsst/L4/GLOB/UKMO/OSTIA/.

    Sea level anomaly (SLA) data are merged products of ERS-1/2, TOPEX/Poseidon, Jason-1/2, and envisat altimeters from the archiving, validation and interpretation of satellite oceanographic data (AVISO) dataset. SLAs represent variations in the sea surface heights relative to the mean sea surface based on a 20-year (1993–2012) reference period. The product has daily records and 0.25°×0.25° resolution. To match the date range of OSTIA SSTs, we used SLA data from 2007 to 2017.

    To identify the ME front, we first identified the ME and SST fronts separately, and then matched the ME and the associated SST front. The data used here are AVISO SLA and OSTIA SST from 2007 to 2017. The flowchart is shown in Fig. 1.

    Figure  1.  Case study of ME front detection on May 30, 2015. Sea level anomaly (a), sea surface temperature (b), SLA of anticyclone eddy (colors) (c), anticyclone (white line) and SST gradient magnitude (colors) (d), and mesoscale eddy (black lines) and mesoscale eddy front (colors) (e). The gray lines in c are isobaths.

    To identify the MEs, we used a winding-angle (WA) algorithm, which detects eddies with greater efficiency, recognition accuracy, and stability (Sadarjoen and Post, 2000; Chaigneau et al., 2008). The WA algorithm searches for local maximum (minimum) SLA values that correspond to potential centers of anticyclonic (cyclonic) eddies in every 4°×4° grid window. Closed streamlines that belong to the same eddy are then selected and clustered by calculating the streamline WA. This algorithm is described in greater detail in Sadarjoen and Post (2000). We used this automatic method to identify MEs in 2007–2017 and obtained 93 anticyclones and 67 cyclones. Then we produced a data array of ME parameters, including ME center, radius, EKE, and shear stress. The identified mesoscale eddy on May 30, 2015 is shown in Fig. 1c.

    We used the maximum gradient magnitude method to detect the SST fronts. Previous studies have used this method to detect SST fronts in the NSCS (Wang et al., 2001; Qiu et al., 2017a). First, we calculated the SST gradient magnitude for pixels (x, y, t) in eight directions from 0° to 360° at 45° intervals, and then we obtained the maximum SST gradient magnitude among the eight directions as GMmax(x, y, t). One case of SST gradient magnitude map is shown in Fig. 1d. Pixels for which GMmax(x, y, t)≥0.2°C/km were identified as frontal pixels and defined as $\, f\left( x,y,t \right)$=1; others were defined as $ \,f\left( x,y,t \right)$=0.

    To determine the ME front position, we defined a search radius, $ r $, for each mesoscale eddy. Then, we assumed that the eddy center located at position $ ({x}_{0},{y}_{0}) $ at time $ t $ and the radius of the mesoscale eddy is $ {r}_{0} $. Thus, the distance between an arbitrary position $ r\left( x,y,t \right)$ and eddy center $ ({x}_{0},{y}_{0}) $ is,

    $$ r^2\left( x,y,t \right)={{\left( x-{{x}_{0}} \right)}^{2}}+{{\left( y-{{y}_{0}} \right)}^{2}},\quad t=1,2,3,\cdots, N. $$ (1)

    $ {\rm{ME}}\left(x,y,t\right) $ is defined as the mesoscale eddy index. Within the search radius of $ \dfrac{3}{2}{r}_{0}({x}_{0},{y}_{0},t) $, we assumed all the pixels are ME pixels, and denoted them as $ {\rm{ME}}\left(x,y,t\right)\!=\!1 $. Beyond this range, $ {\rm{ME}}\left(x,y,t\right)=0 $. That is the pixel $ \left(x,y,t\right) $ with $ {\rm{ME}}\left(x,y,t\right)=1 $ is an ME pixel.

    Then, we matched the MEs and SST fronts. Under the conditions of $ \dfrac{{3r}_{0}({x}_{0},{y}_{0},t)}{2}\!>\! r(x,y,t)\!>\!\dfrac{{r}_{0}({x}_{0},{y}_{0},t)}{2} $, $ {\rm{ME}}\left(x,y,t\right)\!=\! 1 $, and $ \, f\left(x,y,t\right)\!=\!1 $, pixel $ (x,y,t) $ is taken as an ME frontal pixel and marked as $ {\rm{MEF}}\left(x,y,t\right)\!=\!1 $; otherwise, $ {\rm{MEF}}\left(x,y,t\right)\!=\!0. $

    Finally, the absolute probability of the ME front, $ P\left(x,y\right) $, is defined as the count of ME front divided by the sample numbers,

    $$ P\left(x,y\right)=\frac{{\displaystyle\sum\limits _{t=1}^{N}}{\rm{MEF}}\left(x,y,t\right)}{N}. $$ (2)

    The relative probability of the ME front is defined as the ratio between the number of ME fronts and that of mesoscale eddies,

    $$ {P}_{{\rm{r}}}\left(x,y\right)=\frac{{\displaystyle\sum\limits _{t=1}^{N}}{\rm{MEF}}\left(x,y,t\right)}{{\displaystyle\sum\limits _{t=1}^{N}}{\rm{ME}}\left(x,y,t\right)}. $$ (3)

    Although the spatial resolution of satellite SSTs (~5 km) is too coarse to investigate the fine structures of the ME front, it can indicate the presence or absence of the ME front. Hosoda et al. (2012) have attempted to examine multiple scales of oceanic fronts using microwave SSTs with spatial resolutions of 25 km. It is feasible to detect the presence of an ME front using OSTIA SST products.

    We examined a warm eddy from April 19 to June 15, 2015. The SST gradient magnitude map shows that the ME front appeared at the interface between a warm eddy and a cold eddy (Fig. 1e). The three-dimensional structure and track of this eddy were previously reported in Qiu et al. (2019b). The SLA and SST anomaly maps of the eddy on May 30 are shown in Fig. 2a. The SST gradient magnitude along the glider track exceeded 0.2°C/km, reaching the SST front threshold (Fig. 2b). The length of this SST front is approximately 200 km.

    Figure  2.  Sea level anomalies (colors), sea surface temperature anomalies (black lines with interval of 0.25°C) on May 30, 2015 (a), SST gradient magnitude (b), contours of temperature (c), vertical velocity within surface 300 m (d), and vertical diffusion coefficient along the glider track (e). The white line indicates the Chinese underwater glider track. The black dashed box represents the front zone, where the SST GM>0.2°C/km. The vertical diffusion coefficient was calculated by the GHP method.

    We calculated vertical velocities to identify the current across front. Since the geostrophic balance equation may be unsuitable for the ME front, we obtained the vertical velocity w using the density conservation equation (Yu et al., 2019) as follows:

    $$ \frac{\partial \rho }{\partial t}+u\frac{\partial \rho }{\partial x}+v\frac{\partial \rho }{\partial y}+w\frac{\partial \rho }{\partial z}=0, $$ (4)
    $$ w=-\left(\frac{\partial \rho }{\partial t}+u\frac{\partial \rho }{\partial x}+v\frac{\partial \rho }{\partial y}\right)\bigg/\frac{\partial \rho }{\partial z}, $$ (5)

    where ρ is the water density, as observed from Chinese underwater glider data; u and v are from the HYCOM data. We matched the HYCOM data with underwater glider data using a match-up window of 0.1° × 0.1°.

    Vertical mean velocities at 200–800 m are shown in Fig. 2d. The alternation of upward (positive) and downward (negative) velocities within short distances (~10 km) indicates the appearance of complex current structures. Secondary circulation structures can develop vertically in the form of upwelling (downwelling) on the warmer (colder) side of the front (Capet et al., 2008). Our results show more complicated vertical circulation structures. The maximum vertical velocity was approximately 5×10–6 m/s (~4.3 m/d), which was equivalent to that in the northeastern Atlantic (Yu et al., 2019). These vertical velocities could lead to upward or downward heat transport.

    ME fronts may enhance the turbulent kinetic dissipation rate. Therefore, we calculated vertical mixing using Gregg-Henyey-Polzon (GHP) fine-scale parameterization (Henyey et al., 1986; Polzin et al., 2014; Gregg et al., 2003). This method has been successfully used in the SCS (Shang et al., 2017; Liang et al., 2017). The parameterization depends on fine-scale shear and strain variance, as follows:

    $$ K={{K}_{0}}\frac{{{\left\langle V_{z}^{2} \right\rangle }^{2}}}{{\rm{GM}}{{\left\langle V_{z}^{2} \right\rangle }^{2}}}{{h}_{1}}\left( {{R}_{w}} \right)j\left( \frac{f}{N} \right), $$ (6)

    where $ {K}_{0} $ is the reference dissipation rate ($4.0\!\times\! {10}^{-10}$ W/kg), $ {\rm{GM}}\left\langle {V_z^2} \right\rangle $ is the shear variance from the GM model, and $ \left\langle {V_z^2} \right\rangle $ is the observed shear variance.

    $$ {h_1}\left( {{R_w}} \right) = \frac{{3\left( {{R_w} + 1} \right)}}{{2\sqrt 2 {R_w}\sqrt {{R_w} - 1} }}, $$ (7)
    $$ j\left( {\frac{f}{N}} \right) = \frac{{f{\rm{arcosh}}\left( {N/f} \right)}}{{{f_{30}}{\rm{arcosh}}\left( {{N_0}/{f_{30}}} \right)}}, $$ (8)

    where $\,{f_{30}} \!= \!f$(30°N) and ${N}_{0}\!=\!3$ cph. ${R_w} \!= \!\dfrac{{\left\langle {V_z^2} \right\rangle }}{{{{\overline N}^2}\left\langle {\xi _z^2} \right\rangle }}$ is the shear/strain variance ratio, which is approximately 7. The strain variance $\left\langle {\xi _z^2} \right\rangle $ is estimated from buoyancy frequency, $\left\langle {\xi _z^2} \right\rangle \!= \!\Big\langle {{{({N^2} \!- \!{{\overline N}^2})}^2}\!\!/{{\overline N}^4}} \Big\rangle$. Equation (6) can be determined by the strain variance and $ {R}_{w} $,

    $$ {{K}} = {K_0}\frac{{{{\left\langle {\xi _z^2} \right\rangle }^2}}}{{{\rm{GM}}{{\left\langle {\xi _z^2} \right\rangle }^2}}}{h_2}\left( {{R_w}} \right)j\left( {\frac{f}{N}} \right), $$ (9)
    $$ {h_2}\left( {{R_w}} \right) = \frac{{{R_w}\left( {{R_w} + 1} \right)}}{{6\sqrt 2 \sqrt {{R_w} - 1} }}. $$ (10)

    The turbulent dissipation rate, K, is shown in Fig. 2e. Note that K value was very high at the front zone. Enhanced turbulence has been reported for the Kuroshio Current and California fronts (D’Asaro et al., 2011; Johnston et al., 2011). In this study, K is around 10-4 W/kg, larger than that at the ME periphery value observed by Yang et al. (2017). This difference may be due to limitations in the GHP method, which should be used in internal wave-breaking zones (Polzin et al., 2014). Liu et al. (2017) found that the GHP method was not appropriate for the low-latitude northwestern Pacific zone but could show enhanced vertical mixing at the ME boundary. This means that the spatial variation in K is reliable. Yu et al. (2019) suggested that frontogenesis, not frontolysis, can contribute to high turbulence. We also found that some fronts were not associated with increased turbulence (Qiu et al., 2017a). Thus, the enhanced turbulence at the ME front may be induced by frontogenesis.

    Using the ME front detection method described in Section 2.2, we obtained ME front probabilities (Fig. 3). Figures 3a and b show the relative probabilities of the anticyclonic front (AEF) and cyclonic front (CEF). Both AEFs and CEFs had high frequencies (~20%) in the western Luzon Strait and off eastern Hainan Island; thus, approximately five MEs are overlapped with one ME front. The absolute probabilities of AEFs and CEFs were less than 6% in most regions (Figs 3c and d). The highest frequencies of both AEFs and CEFs were observed in the Luzon Strait, which had the highest EKE in the SCS (Zhuang et al., 2010). AEFs were more frequent than CEFs, especially in the western boundary of the SCS. More AEFs were found in the northern part (>20°N) than in the southern part of the NSCS. The total number of ME fronts was high (>150 times) along the western boundary (Figs 3e and f), where MEs intruded the continental slope, and easily produced submesoscale structures and a high turbulent kinetic dissipation rate (Zhang et al., 2016; Yang et al., 2019; Su et al., 2020).

    Figure  3.  Relative probability of ME fronts for anticyclones (a), relative probability of ME fronts for cyclones (b), absolute probability of ME fronts for anticyclones (c), absolute probability of ME fronts for cyclones (d), anticyclonic front number (e), and cyclonic front number (f).

    Mesoscale eddies are asymmetric, so the ME front should have directions around the eddy center. Taking $ \left({x}_{0},{y}_{0}\right) $ as the eddy center, the radial frontal direction can be given by the following equation:

    $$ \left\{ {\begin{aligned} & {A\left( {{x_0},{y_0}} \right) = {\rm{arctan}}\frac{{\left( {y - {y_0}} \right)}}{{\left( {x - {x_0}} \right)}},\qquad\;\;\;\;\;\;\;\;\;\;\;\;\;x > {x_0},{\rm{}}y > {y_0};}\\ & {A\left( {{x_0},{y_0}} \right) = 2{{\text π} } + {\rm{arctan}}\frac{{\left( {x - {x_0}} \right)}}{{\left( {y - {y_0}} \right)}},\;\;\;\;\;\;\;\;\;\;\;\;x > {x_0},{\rm{}}y < {y_0};}\\ & {A\left( {{x_0},{y_0}} \right) = {{\text π} } + {\rm{arctan}}\frac{{\left( {y - {y_0}} \right)}}{{\left( {x - {x_0}} \right)}},\;\;\;\;\;\;\;\;\;\;\;\;\;\;x < {x_0};} \end{aligned}} \right. $$ (11)

    where $ \left({x}_{0},{y}_{0}\right) $ is the center position of the mesoscale eddies and $ \left(x,y\right) $ is the position of the ME front.

    Then, we calculated the probability of AEF and CEF angles at intervals of π/4. The probability maps for AEF and CEF angles are shown in Fig. 4. More than 60% of fronts occurred in the northeastern and southwestern parts of anticyclonic MEs, and less than 40% of fronts occurred in the other two directions. CEFs were almost equally distributed around the cyclonic eddy center, with ~12% in each direction. These results indicate that anticyclones are more asymmetrical than cyclones.

    Figure  4.  The probability of ME front directions for anticyclones (a) and cyclones (b) every π/4.

    Seasonal variations in AEFs and CEFs numbers are shown in Fig. 5. The number of CEFs has a significant seasonal variation, with a maximum value in February (6×105 pixels) and minimum value in September (~0). The number of CEFs exhibited the same seasonal trend as that of generated cyclones, which was suggested to be induced by wind stress curl offshore Luzon Strait and vorticity advection from the Kuroshio Current (Wang et al., 2008; Nan et al., 2011). Seasonal variation in AEFs was not apparent, although AEFs were slightly more frequent from November to March. This is different from the variation of anticyclones, which have significant seasonal variation with larger/smaller numbers in summer/winter due to monsoon-driven positive/negative wind stress curl (Chen et al., 2009). The smaller number of AEFs might result from the decrease in frontal numbers in summer, when solar radiation is almost uniform in the NSCS.

    Figure  5.  Monthly variations in the number of ME fronts (a) and EKE at ME fronts (b). CEFs and AEFs are represented in gray and white colors, respectively.

    To investigate the dynamic process of ME fronts, we calculated the surface eddy kinetic energy (EKE) using ${\rm{EKE}}\!=\!({{u'}}^{2}\!+\!{{v'}}^{2})/2$, where ${u'}\!=\!-\dfrac{g}{f}\dfrac{\partial \left({\rm{SLA}}\right)}{\partial y}$, and ${v'}\!=\!-\dfrac{g}{f}\dfrac{\partial \left({\rm{SLA}}\right)}{\partial x}$ are geostrophic velocity anomalies. $ g $ is the acceleration due to gravity, and $\, f $ is the Coriolis parameter. The seasonal variations in the mean EKEs are shown in Fig. 5b. There is no remarkable seasonal variation in EKE for AEFs and CEFs. Although the number of AEFs is small, the mean EKE in AEFs was quite large in summer. It is clear that the EKE in AEFs was larger than that in CEFs in all the months. The baroclinic front in an AEF is more unstable than that in an CEF (Brannigan, 2016), which may induce EKE to be stronger in AEFs than in CEFs.

    Interannual variations in the numbers of AEFs and CEFs are shown in Fig. 6. The frontal number increased from 2007 to 2010 and then decreased from 2010 to 2018. We compared the numbers of ME fronts with the Niño 3.4 index (Niño 3.4 index; SST anomaly at 5°S to 5°N, 170°W to 120°W). The number of AEF/CEF and ENSO index were not significantly correlated, with coefficient values of –0.3/–0.23. This result is consistent with that revealed by Tuo et al. (2019), who found that the El Niño index and the number of MEs were not significantly correlated after 2004.

    Figure  6.  Time series of numbers of anticyclone fronts (red) and cyclone fronts (blue) (a), Niño3.4 index (b), EKE at cyclone front (c), and EKE at anticyclone front (d). The black and gray lines in b are the annual mean and monthly ENSO index, respectively. The red lines in c and d are the mean EKE magnitudes at the eddy front zone, and black lines are the mean EKE magnitudes within the eddy.

    The total EKE at the ME front and within the eddy are shown in Figs 6c and d. The EKE magnitudes for CEF/AEF have the same trends as those for cyclones/anticyclones. That is when the frontal EKE was strong, the eddy EKE was also strong; thus, the frontal EKE magnitude contributed to the strength of the ME. The total EKE magnitudes were 4 500 and 8 000 cm2/s2 for CEFs and AEFs and 2 000 and 2 100 cm2/s2 for cyclones and anticyclones, respectively. The ME frontal mean EKE magnitudes were nearly 3-fold those of the MEs.

    The SST fronts at the ME edge (ME fronts) may include many processes, including horizontal discrepancies in air-sea heat exchange, wind-induced Ekman current advection, coastal current advection, mesoscale eddy straining, geostrophic disturbance and ageostrophic movement (Hoskins, 1974; Stone and Nemet, 1996). The SST tendency can be calculated from the mixed layer slab model (De Ruijter, 1983). The SST tendency in the radial direction is as follows:

    $$ \frac{{\partial {\rm{SST}}}}{{\partial t}} = - \frac{Q}{{\rho {c_p}h}} + {V_r}\frac{{\partial {\rm{SST}}}}{{\partial r}} + w\frac{{\partial T}}{{\partial z}} + R. $$ (12)

    The four terms on the right-hand side are air-sea heat fluxes, advection, entrainment and diffusion terms. Q is the air-sea net heat flux, and h is the mixed layer depth. Vr and w are the radial and vertical velocities, respectively. T is the temperature at the bottom of the mixed layer. ρ and cp are the sea water density and specific heat capacity. The tendency of the SST gradient magnitude at the edge of the mesoscale eddy can be obtained by taking the r derivation for Eq. (12),

    $$ \begin{split} \frac{\partial }{{\partial t}}\left(\frac{{\partial {\rm{SST}}}}{{\partial r}}\right) =& - \frac{1}{{\rho {c_p}}}\frac{\partial }{{\partial r}}\left(\frac{Q}{h}\right) + \frac{{\partial {V_r}}}{{\partial r}}\frac{{\partial {\rm{SST}}}}{{\partial r}} + \\ &{V_r}\frac{\partial }{{\partial r}}\left(\frac{{\partial {\rm{SST}}}}{{\partial r}}\right) + \frac{\partial }{{\partial r}}\left(w\frac{{\partial T}}{{\partial z}}\right) + \frac{\partial }{{\partial r}}(R). \end{split} $$ (13)

    The air-sea heat flux term in Eq. (13) is difficult to quantify due to the lack of high spatial resolution air-sea net flux products at the eddy edge. The radial velocity Vr is the ageostrophic current, which can improve the water exchange at the eddy edge (Su et al., 2018; Yang et al., 2019), and strengthen or weaken the SST front through advection processes.

    Mesoscale eddies have been proven to modify SST fronts in previous studies. Eddy straining is a frontogenesis mechanisms and has submesoscale characteristics (Capet et al., 2008; Brannigan, 2016). Eddy straining or deformation and asymmetry are suggested to be influenced by background large-scale currents, i.e., wind-driven currents and Kuroshio Current (Qiu et al., 2019b). In the NSCS, submesoscale structures are also active (Yang et al., 2017; Qiu et al., 2019a; Zheng et al., 2008). Dong and Zhong (2018) examined the spatiotemporal features of submesoscale processes and found that these processes are strong in winter and weak in summer. Their generation mechanisms have been revealed by Zhang et al. (2020).

    SST fronts in the NSCS have both mesoscale/large-scale (Re<1) and submesoscale (Re~O(1)) structures (Wang et al., 2001; Jing et al., 2015; Zhong et al., 2017). In our case study (Fig. 2), the Rossby number of the SST front is Re<1 with a spatial range of 200 km and a mean horizontal velocity of 0.5 m/s. Therefore, this front is a meso/large scale front. However, at the warm and cold eddy edge, two narrow maximum vertical velocity zones with widths of ~10 km occurred (Fig. 2d), which are characteristic of submesoscale structures (Re~O(1)). This indicates that a large-scale SST front also includes submesoscale structures at the mesoscale eddy edge. To quantify the contributions of mesoscale eddies on the SST fronts, high spatial resolution observational data sets, including air-sea heat fluxes and oceanic physical parameters (temperature, salinity and velocity), are needed in the future studies.

    Using satellite data, we developed an automatic integrated method to detect ME SST fronts in the NSCS and examined their spatiotemporal variations. ME fronts occupied 20% of the MEs; the northeast and southwest parts of anticyclones were more prone to generating fronts. Mean EKE values at the ME fronts were three times those of the MEs. CEFs were more common from winter to spring, and AEFs were common in all months. The EKE in an AEF was larger than that in an CEF, which might be due to the different levels of AEFs and CEFs.

    The results of the current study can be used as a benchmark for future in situ ME front observations. We only detected SST fronts overlapping at the edge of the ME, which may be one part of a large-scale SST front, and the mesoscale eddy only modulates one part of the large SST front. We need to quantify the mesoscale eddy contributions on ME fronts by using high spatiotemporal resolution in situ data in the future.

    Chinese underwater glider data were provided by State Key Laboratory of Robotics, Shenyang Institute of Automation, Chinese Academy of Science. We thank the AVISO (http://www.aviso.oceanobs.com/en/data/products/sea-surface-height-products/global/index.html) for sea level anomaly data, HYCOM for current velocity data, and OSTIA (http://ghrsst-pp.metoffice.com/pages/latest_analysis/ostia.html) for SST data.

  • Aki K, Lee W H K. 1976. Determination of three-dimensional velocity anomalies under a seismic array using first P arrival times from local earthquakes: 1. A homogeneous initial model. Journal of Geophysical Research, 81(23): 4381–4399. doi: 10.1029/JB081i023p04381
    Cai Laixing, Xiao Guolin, Guo Xingwei, et al. 2019. Assessment of Mesozoic and Upper Paleozoic source rocks in the South Yellow Sea Basin based on the continuous borehole CSDP-2. Marine and Petroleum Geology, 101: 30–42. doi: 10.1016/j.marpetgeo.2018.11.028
    Gao Xiaohui, Zhang Xunhua, Guo Xingwei, et al. 2020. Provenance and tectonic implications of Paleozoic strata in the South Yellow Sea Basin, China-revealed from the borehole CSDP-2. Journal of Ocean University of China, 19(3): 536–550. doi: 10.1007/s11802-020-4088-y
    Hao Tianyao, Huang Song, Xu Ya, et al. 2010. Geophysical understandings on deep structure in Yellow Sea. Chinese Journal of Geophysics (in Chinese), 53(6): 1315–1326
    Hao Tianyao, Suh M, Wang Qianshen, et al. 2002. A study on the extension of fault zones in Yellow Sea and its adjacent areas based on gravity data. Chinese Journal of Geophysics (in Chinese), 45(3): 385–397
    He Enyuan, Zhao Minghui, Qiu Xuelin, et al. 2016. Crustal structure across the post-spreading magmatic ridge of the East Sub-basin in the South China Sea: tectonic significance. Journal of Asian Earth Sciences, 121: 139–152. doi: 10.1016/j.jseaes.2016.03.003
    Hong T K, Choi H. 2012. Seismological constraints on the collision belt between the North and South China blocks in the Yellow Sea. Tectonophysics, 570–571: 102–113
    Hou Hesheng, Gao Rui, Lu Zhanwu, et al. 2009. First arrival seismic tomographic imaging test of the near-surface velocity structure of central uplift in the Qiangtang Basin, Qinghai-Tibet Plateau. Geological Bulletin of China (in Chinese), 28(6): 738–745
    Huang Song, Hao Tianyao, Xu Ya, et al. 2010. Study on macro distribution of residual basin of South Yellow Sea. Chinese Journal of Geophysics (in Chinese), 53(6): 1344–1353
    Kim H J, Kim C H, Hao Tianyao, et al. 2019. Crustal structure of the Gunsan Basin in the SE Yellow Sea from ocean bottom seismometer (OBS) data and its linkage to the South China Block. Journal of Asian Earth Sciences, 180: 103881. doi: 10.1016/j.jseaes.2019.103881
    Lei Baohua, Chen Jianwen, Liang Jie, et al. 2018a. Tectonic deformation and evolution of the South Yellow Sea basin since Indosinian movement. Marine Geology & Quaternary Geology (in Chinese), 38(3): 45–54
    Lei Baohua, Chen Jianwen, Wu Zhiqiang, et al. 2018b. Density and velocity analysis and seismic reflection model construction of marine Mesozoic-Paleozoic in the North Jiangsu-to-South Yellow Sea Basin. Oil Geophysical Prospecting (in Chinese), 53(3): 558–576
    Li Sanzhong, Jahn B M, Zhao Shujuan, et al. 2017. Triassic southeastward subduction of North China Block to South China Block: insights from new geological, geophysical and geochemical data. Earth-Science Reviews, 166: 270–285. doi: 10.1016/j.earscirev.2017.01.009
    Li Nan, Li Weiran, Long Haiyan, 2016. Tectonic evolution of the north depression of the South Yellow Sea basin since late Cretaceous. Journal of Ocean University of China, 15(6): 967–976
    Liu Kai, Liu Huaishan, Wu Zhiqiang, et al. 2016. Seismic acquisition parameters analysis for deep weak reflectors in the South Yellow Sea. Journal of Ocean University of China, 15(5): 758–766. doi: 10.1007/s11802-016-2978-9
    Pang Yumao, Guo Xingwei, Han Zuozhen, et al. 2019. Mesozoic–Cenozoic denudation and thermal history in the Central Uplift of the South Yellow Sea basin and the implications for hydrocarbon systems: constraints from the CSDP-2 borehole. Marine and Petroleum Geology, 99: 355–369. doi: 10.1016/j.marpetgeo.2018.10.027
    Qi Jianghao, Wu Zhiqiang, Guo Xingwei, et al. 2019. Application of large capacity air gun source in VSP data acquisition under marine high speed shielding layer in South Yellow Sea: illustrated by the example of well CSDP (Continental Shelf Drilling Program)-2. Progress in Geophysics (in Chinese), 34(4): 1661–1670
    Shinn Y J. 2015. Geological structures and controls on half-graben inversion in the western Gunsan Basin, Yellow Sea. Marine and Petroleum Geology, 68: 480–491. doi: 10.1016/j.marpetgeo.2015.09.013
    Shinn Y J, Chough S K, Hwang I G. 2010. Structural development and tectonic evolution of Gunsan Basin (Cretaceous-Tertiary) in the central Yellow Sea. Marine and Petroleum Geology, 27(2): 500–514. doi: 10.1016/j.marpetgeo.2009.11.001
    Tao Kai, Grand S P, Niu Fenglin. 2018. Seismic structure of the upper mantle beneath eastern asia from full waveform seismic tomography. Geochemistry, Geophysics, Geosystems, 19(8): 2732–2763,
    Wang Wei, Chen Gao, Wang Jialin, et al. 1999. Analysis for regional structural characteristics of North Jiangsu-South Yellow Sea basin. Journal of Seismology (in Chinese), (1): 47–55
    Wang Jian, Zhao Minghui, He Enyuan, et al. 2014. The selection of optimal inversion parameters for first-arrival seismic tomography: an application to 3D seismic data from the central sub-basin of the South China Sea. Journal of Tropical Oceanography (in Chinese), 33(5): 74–83
    Weekly R T, Wilcock W S D, Toomey D R, et al. 2014. Upper crustal seismic structure of the Endeavour segment, Juan de Fuca Ridge from traveltime tomography: implications for oceanic crustal accretion. Geochemistry, Geophysics, Geosystems, 15(4): 1296–1315
    Wu Zhiqiang. 2009. The seismic techniques for exploring marine facies stratigraphic hydrocarbon entrapped in the Middle Uplift of the South Yellow Sea (in Chinese)[dissertation]. Qingdao: Ocean University of China
    Wu Zhiqiang, Liu Lihua, Xiao Guolin, et al. 2015. Progress and enlightenment of integrated geophysics exploration of marine residual basin in the South Yellow Sea. Progress in Geophysics (in Chinese), 30(6): 2945–2954
    Wu Zhiqiang, Qi Jianghao, Zhang Xunhua, et al. 2019. Vertical seismic profiling survey on the Well CSDP-2 of the “continental shelf drilling program”. Chinese Journal of Geophysics (in Chinese), 62(9): 3492–3508
    Yang Jinyu. 2009. Research on the tectonic relation between the South Yellow Sea basin and its adjacent area and distribution characteristic and tectonic evolution of the Mesozoic-Paleozoic marine strata (in Chinese)[dissertation]. Hangzhou: Zhejiang University
    Yang Yanqiu, Li Gang, Yi Chunyan. 2015. Characteristics of seismic reflection and geological ages of seismic sequences for marine strata in the South Yellow Sea basin. Journal of Northeast Petroleum University (in Chinese), 39(3): 50–59
    Yao Yongjian, Feng Zhiqiang, Hao Tianyao, et al. 2008. A new understanding of the structural layers in the South Yellow Sea Basin and their hydrocarbon-bearing characteristics. Earth Science Frontiers, 15(6): 232–240
    Yi S, Yi S, Batten D J, et al. 2003. Cretaceous and Cenozoic non-marine deposits of the northern South Yellow Sea Basin, offshore western Korea: palynostratigraphy and palaeoenvironments. Palaeogeography, Palaeoclimatology, Palaeoecology, 191(1): 15–44
    Yuan Yong, Chen Jianwen, Zhang Yinguo, et al. 2018. Sedimentary system characteristics and depositional filling model of Upper Permian–Lower Triassic in South Yellow Sea Basin. Journal of Central South University, 25(12): 2910–2928. doi: 10.1007/s11771-018-3962-x
    Zelt C A, Barton P J. 1998. Three-dimensional seismic refraction tomography: a comparison of two methods applied to data from the Faeroe Basin. Journal of Geophysical Research: Solid Earth, 103(B4): 7187–7210. doi: 10.1029/97JB03536
    Zhang Haiqi, Chen Jianwen, Li Gang, et al. 2009. Discovery from seismic survey in Laoshan Uplift of the South Yellow Sea and the significance. Marine Geology & Quaternary Geology (in Chinese), 29(3): 107–113
    Zhang Xunhua, Wu Zhiqiang, Liu Lihua, et al. 2022. Submarine Seismic Exploration in the Eastern China Sea (in Chinese). Beijing: Science Press
    Zhang Minghua, Xu Deshu, Chen Jianwen. 2007. Geological structure of the Yellow Sea Area from regional gravity and magnetic interpretation. Applied Geophysics, 4(2): 75–83. doi: 10.1007/s11770-007-0011-1
    Zhang Xunhua, Yang Jinyu, Li Gang, et al. 2014. Basement structure and distribution of Mesozoic-Paleozoic marine strata in the South Yellow Sea basin. Chinese Journal of Geophysics (in Chinese), 57(12): 4041–4051
    Zhang Xunhua, Zhang Zhixun, Lan Xianhong, et al. 2013. Regional Geology in South Yellow Sea (in Chinese). Beijing: China Ocean Press
    Zhang Xiaohua, Zhang Xunhua, Wu Zhiqiang, et al. 2018. New understanding of Mesozoic-Paleozoic strata in the Central Uplift of the South Yellow Sea basin from the drilling of well CSDP-02 of the “Continental Shelf Drilling Program”. Chinese Journal of Geophysics (in Chinese), 61(6): 2369–2379
    Zhao Minghui, He Enyuan, Sibuet J C, et al. 2018. Postseafloor spreading volcanism in the central east South China Sea and its formation through an extremely thin oceanic crust. Geochemistry, Geophysics, Geosystems, 19(3): 621-641
    Zhao Weina, Wang Huigang, Shi Hongcai, et al. 2019a. Crustal structure from onshore-offshore wide-angle seismic data: application to northern Sulu Orogen and its adjacent area. Tectonophysics, 770: 228220. doi: 10.1016/j.tecto.2019.228220
    Zhao Weina, Zhang Xunhua, Wang Huigang, et al. 2020. Characteristics and noise combination suppression of wide-angle Ocean Bottom Seismography (OBS) data in shallow water: a case study of profile OBS2016 in the South Yellow Sea. Chinese Journal of Geophysics (in Chinese), 63(6): 2415–2433
    Zhao Weina, Zhang Xunhua, Zou Zhihui, et al. 2019b. Velocity structure of sedimentary formation in the South Yellow Sea Basin based on OBS data. Chinese Journal of Geophysics (in Chinese), 62(1): 183–196
    Zou Zhihui, Liu Kai, Zhao Weina, et al. 2016. Upper crustal structure beneath the northern South Yellow Sea revealed by wide-angle seismic tomography and joint interpretation of geophysical data. Geological Journal, 51(S1): 108–122. doi: 10.1002/gj.2847
  • Relative Articles

  • Cited by

    Periodical cited type(1)

    1. Juan Gabriel Rueda-Bayona, José Luis García Vélez, Daniel Mateo Parrado-Vallejo. Effect of Sea Level Rise and Access Channel Deepening on Future Tidal Power Plants in Buenaventura Colombia. Infrastructures, 2023, 8(3): 51. doi:10.3390/infrastructures8030051

    Other cited types(0)

  • 加载中

Catalog

    通讯作者: 陈斌, bchen63@163.com
    • 1. 

      沈阳化工大学材料科学与工程学院 沈阳 110142

    1. 本站搜索
    2. 百度学术搜索
    3. 万方数据库搜索
    4. CNKI搜索

    Figures(15)  / Tables(1)

    Article Metrics

    Article views (813) PDF downloads(50) Cited by(1)
    Proportional views
    Related

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return