Abstract
The timing of the onset of Taiwan sediment input to the northern South China Sea (SCS) is still controversial even though many provenance proxies had been used. To analyze the change of Taiwan input since the Late Miocene (11.63–0 Ma), we applied the major-element-based proxies R AK and R KCN, combined with the published clay mineral and Sr–Nd isotope data, to study the sediment provenance near the northern SCS slope. The results show that significant Taiwan sediment input began at ∼6.5 Ma in the Late Miocene, consistent with the timing of Taiwan uplift. Moreover, Pearl River input increased after ∼3.6 Ma, corresponding to the sea level fall caused by the ice sheet expansion in the Northern Hemisphere. The Taiwan input increased during the Mid-Pleistocene Transition (MPT, 1.25–0.7 Ma) because of the Northern Pacific Intermediate Water expansion. After the MPT, the Pearl River input re-increased in response to the lowered sea level in glacials. In general, tectonic activities such as the Taiwan uplift control the sediment provenance in the northern SCS from 6.5 to 3.6 Ma. After 3.6 Ma, as the weathering regime of Taiwan sediment became more stable, climate change became a more important factor in influencing sediment provenance.
1 Introduction
Variations of sedimentary sources and transport processes will profoundly affect the composition of deep-sea sediments [1]. Marginal seas represent a critical component of the global source-to-sink system and they are ideal sites to investigate land–sea interactions, because the sediment budget and transport pathways can be more easily defined quantitatively in these areas [2,3]. As the rivers in coastal region of SE-Asia provide the largest portion of global sediment discharge to the ocean [4], the South China Sea (SCS) offers an excellent case for the source-to-sink study among the global marginal seas [5].
In previous studies, the sedimentation rates, clay minerals, trace elements, and Sr–Nd isotopes have been used to trace the onset timing of the Taiwan input, and the results mainly range from 5 to 3 Ma [6–8]. However, the uplift of the primordial Taiwan Island began at 6.5 Ma in the Late Miocene [9,10], and the denudation process began immediately after the uplift [11]. Taiwan-sourced sediments were also found in the foreland and forearc basins around the Taiwan island at ∼6.5 Ma [12–15]. Moreover, the water mass that transports Taiwan’s materials to the northern deep water area of the SCS, i.e., the North Pacific Deep Water (NPDW), started to enter the SCS at 23.8 Ma [16,17]. The South China Sea Brunch of Kuroshio (SCSBK) and Northern Pacific Intermediate Water (NPIW) could also be the possible pathways of the Taiwan sediment since the Late Miocene due to their westward flow direction [18,19]. Thus, the necessary conditions for the denudation and transportation of Taiwan-sourced sediments were already in existence at 6.5 Ma, and Taiwan input should have been seen in the northern SCS since the Late Miocene. The seismic data and multibeam bathymetry proved that more Taiwan sediment was transported to the northern SCS basin by intensified bottom currents since ∼6.5 Ma [20].
It is difficult to apply conventional sediment provenance proxies to the northern SCS, since the radiogenic Nd isotope (ε Nd) values of the Taiwan and the Pearl River inputs are too close to be distinguished [1], and the clay minerals are more likely to indicate changes in fine-grained, rather than bulk, sediment provenance [21]. Therefore, a proxy that can not only distinguish the Pearl River and Taiwan sources, but also represent the bulk sediment could be more effective to trace the variations of sediment provenance in the northern SCS.
The recently developed major element-based provenance proxies (R AK and R KCN) [22] meet these two requirements. The chemical weathering in the Pearl River basin was much stronger than Taiwan today [1], and the weathering intensity of Taiwan sediment was always insufficient since 6.5 Ma because of the strong physical erosion [9,23,24]. The R AK and R KCN are the proxies of the weathering trend, which can be analyzed by correlation coefficients between the chemical index of alteration (CIA), K*, and CN*. The CIA is defined as Al2O3/(Al2O3 + CaO* + Na2O + K2O) × 100, where CaO* represents the CaO in the silicate fraction [25], the correction for phosphate phases is straightforward by CaO* = CaO−3.33P2O5 assuming that all P2O5 is associated with apatite [26,27]. The K* is defined as K2O/(Al2O3 + CaO* + Na2O + K2O) × 100, the CN* is defined as (CaO* + Na2O)/(Al2O3 + CaO* + Na2O + K2O) × 100 [22]. The R AK is the Pearson’s r between CIA and K*, and R KCN is the Pearson’s r between K* and CN* [22]. Here we use the data of IODP Site U1501 and ODP Sites 1144, 1146, and 1148 to calculate the changes of R AK and R KCN since the Late Miocene. Together with the data of oxygen isotopes, Sr–Nd isotopes, clay minerals, and grain size, we analyzed the changes of provenance assemblages in the northern SCS since the Late Miocene (11.63 Ma). We also reassess the influence of tectonics and climate change on the sediment composition of the northern SCS.
2 Regional setting
The SCS is the largest marginal sea in the western Pacific, while the latitude range spans from 0° to 23°N, the wind and precipitation patterns are dominated by the East Asian Monsoon [1]. Given the small flux of aeolian dust, most of the sediments in the northern SCS are transported by the fluvial system [1,28]. Since the Quaternary, the Pearl River and Taiwan have become the principal sediment sources to the northern SCS (Figure 1) [34,35]. The relative supply of Taiwan and the Pearl River sediment is mainly controlled by the ocean current pattern and climate change [36].
![Figure 1
Bathymetric map and the location of sites in the SCS and North Pacific. Sites labeled with red-filled circles are studied in this work. Sites labeled with green-filled circles are those with published data. The annual sediment discharges of potential sources are indicated by red arrows [4]. SCSBK: South China Sea Branch of Kuroshio (grey arrow) [19,29]; NPIW: North Pacific Intermediate Water (purple arrow) [18]; SCSIW: South China Sea Intermediate Water [30]; NPDW: North Pacific Deep Water (yellow arrow) [31,32]. EASM: East Asian summer monsoon; and EAWM: East Asian winter monsoon [33].](/document/doi/10.1515/geo-2022-0454/asset/graphic/j_geo-2022-0454_fig_001.jpg)
Bathymetric map and the location of sites in the SCS and North Pacific. Sites labeled with red-filled circles are studied in this work. Sites labeled with green-filled circles are those with published data. The annual sediment discharges of potential sources are indicated by red arrows [4]. SCSBK: South China Sea Branch of Kuroshio (grey arrow) [19,29]; NPIW: North Pacific Intermediate Water (purple arrow) [18]; SCSIW: South China Sea Intermediate Water [30]; NPDW: North Pacific Deep Water (yellow arrow) [31,32]. EASM: East Asian summer monsoon; and EAWM: East Asian winter monsoon [33].
Since the SCS formed in the early Oligocene, its principal fluvial sediment provenance has changed over time. During the Oligocene, the main source of the northern SCS was the central Vietnam [37]. During most of the Miocene, Luzon was still far to the south, and Taiwan had not yet significantly risen above the sea surface, while the Pearl River became the primary source [21,37,38]. In consideration of the two end members including the Pearl River and Taiwan in modern times, the onset of Taiwan input should be the most important change of sediment provenance in the northern SCS since the Miocene.
The sediment particles are transported both alongslope and downslope to the northern slope by oceanic currents [39]. Along the slope, there are four contour currents at different water depths (Figure 1): the westward SCSBK (<500 m) [40], the westward NPIW (around 300–800 m) [41,42], the eastward South China Sea Intermediate Water (SCSIW, around 500–1,500 m) [30], and the westward NPDW (>2,000 m) [43,44]. In addition, the downslope gravity flows may transport a lot of coarse deposits from shelf to basin around the Pearl River Canyon, formed the mass-transport deposits or turbidites [45,46]. The westward contour currents always bring the particles from Taiwan or Dongsha to the shelf, upper slope, and basin, while the eastward current SCSIW and gravity flows are in favor of the transportation from the South China [22,39]. The modern sediment transport pattern in the SCS could exist at least since the Late Miocene (11.63 Ma) [31,47].
3 Materials and methods
3.1 IODP Site U1501 and samples
We focus on sediments recovered from the continental slope of the northern SCS (Figure 1). IODP Site U1501 (18°53.09′N, 115°45.95′E, water depth: 2,846 m) is situated on the west side of the Pearl River Canyon [48,49] and the Ocean Drilling Program (ODP) Sites 1144 (20°3.18′N, 117°25.14′E, water depth: 2,037 m), 1146 (19°27.40′N, 116°16.37′E, water depth: 2,092 m), and 1148 (18°50.17′N, 116°33.94′E, water depth: 3,294 m) are located on the east side [5]. Thirty-one samples were selected from IODP Site U1501 at 20 cm intervals between 99.29 and 92.86 m, which spans from 5.8 to 5.2 Ma [48]; all samples were analyzed for major elemental concentrations on inductively coupled plasma-optical emission (ICP-OES). Moreover, the major elemental data of the 46 samples of ODP Sites 1144 (1–0 Ma) [23], 202 samples of ODP Site 1146 (11.63–0 Ma) [6], and 52 samples of ODP Site 1148 (11.63–0 Ma) [50] are compiled from the literature.
3.2 ICP-OES analysis
The sediment samples from IODP Site U1501 were ground to 200 mesh. Each sample was dried, weighed with an accuracy 100 mg, placed in a centrifuge tube. Detrital silicates in sediments were extracted following a sequential leaching procedure in four steps [51,52]. First, a CH3COONH4/CH3COOH buffer solution at pH 4.5 was added to remove carbonates, then ultrasonic cleaning was done at 60°C for 3 h. Second, the organic matter was removed by 10% H2O2 at 60°C in an ultrasonic bath for 1 h repeated three times. Third, the silicate residues were dissolved using a concentrated HF + HNO3 mixture on a hot plate at 130°C for 12 h, then in a HNO3 + HCl mixture at 130°C for 12 h. Finally, samples were diluted with 2% HNO3 for major element measurement. The major elements were measured on a Spectro Blue Sop ICP-OES at Peking University. The external reproducibility for the elements (Na, Mg, Al, K, Ca, P, and Ti) in silicate residues was ±5%.
3.3 Sediment provenance proxies
The provenance proxies R AK and R KCN are developed from the CIA based on major element contents, which can indicate the direction of the weathering trend [22]. The CIA is defined as Al2O3/(Al2O3 + CaO* + Na2O + K2O) × 100, and CaO* was replaced by CaO−3.33P2O5 because CaO/Na2O ratio was <1 in all samples [26,27]. The values of R AK and R KCN are calculated for every continuous 10 points [22]. Pearl River sourced sediments typically have negative values of R AK and near-zero values of R KCN due to the more intense chemical weathering and leaching of potassium [1]. In contrast, Taiwan sourced sediments have positive values of R AK and negative values of R KCN, because the physical erosion is stronger relative to chemical weathering, and the potassium-rich clay minerals such as illite are abundant [1].
4 Results
In the A-CN-K triangle diagram (Figure 2), the major element contents of ODP Sites 1146 and 1148 are closer to the Pearl River catchment before 6.5 Ma, but fall clearly within the Taiwan region after 6.5 Ma (Figure 2). In addition, the geochemical space spanned by the sediments from ODP Site 1144 since 1 Ma is also near the Taiwan source range. In contrast, the major element contents of IODP Site U1501 (5.8–5.2 Ma) are almost all within the Pearl River range. In addition, the data of elemental content including Al2O3, K2O, Na2O, CaO, and P2O5 can be accessed in the supplementary data.
![Figure 2
Major element concentrations in the northern SCS drilling sites since the Late Miocene. (a) Major elemental Al2O3–(CaO + Na2O)–K2O(A-CN-K) diagram of bulk and clay-fraction sediments in surrounding fluvial drainage systems [1], IODP Site U1501 (this study), and ODP Sites 1144 [23], 1146 [6], and 1148 [50]. (b) Correlations of the CIA with K2O/(CaO* + Na2O) molar ratio of argillaceous sediments in the Pearl River, Taiwan, Luzon, IODP Site U1501, and ODP Sites 1144, 1146, and 1148.](/document/doi/10.1515/geo-2022-0454/asset/graphic/j_geo-2022-0454_fig_002.jpg)
Major element concentrations in the northern SCS drilling sites since the Late Miocene. (a) Major elemental Al2O3–(CaO + Na2O)–K2O(A-CN-K) diagram of bulk and clay-fraction sediments in surrounding fluvial drainage systems [1], IODP Site U1501 (this study), and ODP Sites 1144 [23], 1146 [6], and 1148 [50]. (b) Correlations of the CIA with K2O/(CaO* + Na2O) molar ratio of argillaceous sediments in the Pearl River, Taiwan, Luzon, IODP Site U1501, and ODP Sites 1144, 1146, and 1148.
We used the element data of IODP Site U1501 and ODP Sites 1144, 1146, and 1148 to calculate the provenance indicators R AK and R KCN. The results showed that the materials of IODP Site U1501 mainly came from the Pearl River during most of the time from 5.7 to 5.3 Ma, except for a brief period that was dominated by Taiwan input around 5.55–5.52 Ma (Figure 3d). Among the other three sites, the sediments of ODP Site 1144 were predominantly Taiwan-derived from 0 to 1 Ma (Figure 3c); to the west ODP Site 1146 shows clear signs of Taiwan input around 6.5, 5.9–5.7, 5.5–5.4, 4.9–4.6, 3.8–3.2, 1.4, and 1.2–0.8 Ma (Figures 3b and 4a); and the deeper ODP Site 1148 exhibits Taiwan-derived input around 6 and 1.3–1 Ma (Figures 3b and 4a).
![Figure 3
Major elemental provenance proxies R
AK and R
KCN values since the Late Miocene from: (a) ODP Site 1146, (b) ODP Site 1148, (c) ODP Site 1144 [22], and (d) IODP Site U1501. The red bars mark the time intervals with significant Taiwan sediment input. The red line highlights the start of Taiwan orogen [9].](/document/doi/10.1515/geo-2022-0454/asset/graphic/j_geo-2022-0454_fig_003.jpg)
Major elemental provenance proxies R AK and R KCN values since the Late Miocene from: (a) ODP Site 1146, (b) ODP Site 1148, (c) ODP Site 1144 [22], and (d) IODP Site U1501. The red bars mark the time intervals with significant Taiwan sediment input. The red line highlights the start of Taiwan orogen [9].
![Figure 4
Temporal evolution in proxies linked to chemical weathering intensity and climate indices since the Late Miocene. (a) R
AK values of IODP Site U1501 and ODP Sites 1144, 1146, and 1148. (b) Global [53] and Pearl River Mouth Basin eustasy sea level curve [54]. (c) Percentages of quartz from IODP Sites U1499, U1501, and U1505 [55] and relative quartz from ODP Sites 1146 and 1148 [38,56]. (d) Mean grain size of bulk sediment from ODP Sites 1146 and 1148 [38,57]. (e) Kaolinite/illite values from ODP Sites 1146 and 1148 [24,58]. (f) δ18OPF and CIA of ODP Site 1146 [6,59,60] and ODP Site 1148 [23,61]. The grey bar highlights the uncoordinated variation interval between the CIA and δ18O values. The red line highlights the start of Taiwan orogen [9].](/document/doi/10.1515/geo-2022-0454/asset/graphic/j_geo-2022-0454_fig_004.jpg)
Temporal evolution in proxies linked to chemical weathering intensity and climate indices since the Late Miocene. (a) R AK values of IODP Site U1501 and ODP Sites 1144, 1146, and 1148. (b) Global [53] and Pearl River Mouth Basin eustasy sea level curve [54]. (c) Percentages of quartz from IODP Sites U1499, U1501, and U1505 [55] and relative quartz from ODP Sites 1146 and 1148 [38,56]. (d) Mean grain size of bulk sediment from ODP Sites 1146 and 1148 [38,57]. (e) Kaolinite/illite values from ODP Sites 1146 and 1148 [24,58]. (f) δ18OPF and CIA of ODP Site 1146 [6,59,60] and ODP Site 1148 [23,61]. The grey bar highlights the uncoordinated variation interval between the CIA and δ18O values. The red line highlights the start of Taiwan orogen [9].
5 Discussion
5.1 Negligible aeolian dust input in the Late Miocene
The chemical weathering intensity is mainly controlled by temperature in the source areas [62]. The precipitation intensity influenced by the East Asian summer monsoon mainly affects the physical denudation and river runoff [62]. The East Asian winter monsoon (EAWM) mainly influences the transport of aeolian particles [38]. In other words, the environmental changes can significantly alter the sediment material and elemental composition.
The planktonic foraminifera oxygen isotope and CIA can be regarded as the proxies of local temperature and chemical weathering intensity, respectively [59–61]. We find that the CIA of ODP Sites 1146 and 1148 is in good agreement with oxygen isotope variability most of the time since the Late Miocene except 6.5–3 Ma (Figure 4f).
Based on previous studies, the decrease of CIA after 6.5 Ma seems mainly due to the increase of grain size and quartz content [24,38]. However, the negative correlations between CIA and mean grain size of ODP Sites 1146 and 1148 are not significant since the Late Miocene (Figure 5). In general, stronger physical and chemical weathering processes produce sediments with finer grain size [63], and an increase in relative quartz content often leads to a decrease in weathering sensitive minerals [64]. The average grain size and quartz content of sediments at ODP Site 1146 around 6.5 Ma are the largest since 11.63 Ma (Figure 4c and d) [38]. As the winter monsoon intensity increased at 8–6 Ma [65] the wind dust particles were also found at ODP Site 1146 during the Late Miocene [38], these coarse-grained materials with a modal grain size of ∼10 µm were therefore thought to be mainly aeolian dust from inland Asia [38]. However, rivers around the modern SCS can also provide large amounts of >10 μm particles, which are widely distributed on the continental slope [39].
We found that the deposition of distal aeolian dust is likely negligible based on a comprehensive comparison of several studies in inland Asia and the Pacific Ocean. Aeolian dust deposited in the SCS is mainly transported from inland Asia by EAWM today. The Chinese Loess Plateau (CLP) has been consistently regarded as the source area of aeolian dust deposited in the SCS. However, the sediment of >10 μm grain from the CLP is mainly derived from neighboring river systems since the Quaternary [66]. The winter wind mainly transported the aeolian dust at the low altitude within the elevation range of 0–3,000 m, and the particles >10 μm would mostly settle within a horizontal distance of 1,000 km [67]. Meanwhile, the particles <10 μm could be transported for thousands of kilometers, likely to the Pacific Ocean and SCS (Figure 1) [68,69]. The transport distance of the aeolian dust from the CLP to the SCS is more than 3,000 km, and the modern average grain size is 5.75 μm during winter and 3.62 μm from summer [70]. We suggest that the aeolian dust flux should be even smaller with finer grain size in the Late Miocene. First, the intensity of EAWM roughly shows a stepwise enhancement since 8 Ma [71,72], thus the winter monsoon may be weaker in the Late Miocene relative to today. Second, the aeolian dust source area in the inland Asia was much smaller than today [73]. In addition, the mean grain size of the aeolian dust in the northern Pacific and Philippine Sea is always less than 5 μm (Figure 1) [74,75], which is much smaller than the dust in the northern SCS [38]. In addition, the high vegetation cover in the South China had a strong inhibitory effect on the formation of aeolian dust [76,77], even at the glacials on the emerged continental shelf [78,79].
The sediment grain size distribution in the northern SCS is spatially different since the Late Miocene. IODP Sites U1499, U1501, and U1505 are located about 100 km west of ODP Sites 1146 and 1148 (Figure 1), and there are no obvious increases in quartz percentage from 6.5 to 3 Ma at these sites (Figure 4c). The CIA values of IODP Site U1501 are also significantly higher than that of ODP Sites 1146 and 1148 (Figure 4f). For long-distance transport over 3,000 km, the particle size, flux, and chemical composition of wind and dust materials should be homogeneous throughout the northern SCS. Therefore, the spatial differences in sediment composition are more likely to be the result of the interaction between near-field deposition and oceanographic current distribution.
5.2 Taiwan input since the Late Miocene
The weakly-weathered sediments in the northern SCS since 6.5 Ma are more likely from Taiwan. At the beginning of the Late Miocene, the Pearl River became the dominant sediment source to the northern SCS [37]. Around 6.5 Ma, the proto-Taiwan began to rise above sea level [9]. The intense tectonic activities in Taiwan since the Late Miocene would likely strengthen the physical erosion and suppress the chemical weathering [1,14,24]. And the Taiwan sediments could have been transported to the northern SCS, since the instruction of NPDW started at the Early Miocene [16,17].
The weakly-weathered sediments cannot be mainly transported by rivers from the South China, because the chemical weathering intensity of the Pearl River basin was likely even greater in a hot humid climatic condition in the past. The chemical weathering of the Pearl River basin is mainly controlled by temperature and precipitation [80]. The temperature in the Late Miocene was even higher (Figure 4f) [6,50,59–61], and the precipitation was also stronger [81,82]. The warm and humid climate resulted in the growth of a large amount of vegetation; meanwhile, the continuously rising sea level [53,54,83] and the subsidence of the SCS basin [34,57,84] in the Late Miocene likely made the exposure of the shelves improbable. However, the 40–50 m variations of sea level still existed in the Late Miocene (Figure 4b), and the shelf erosion by rivers and waves might have played an essential role.
There are obvious differences in the weathering characteristics between the Pearl River and Taiwan. The chemical weathering intensity in Taiwan is lower and the potassium leaching process is much less sufficient than the Pearl River today [1]. The differences in weathering regimes should already exist in the Late Miocene. Statistically speaking, the R AK values of ODP Sites 1146 and 1148 are always negative from 11.63 to 7 Ma (Figure 6a), indicating that most of the materials at these two sites come from the Pearl River basin at this time, which is consistent with the results of Sr–Nd isotopes, La/Th, and other provenance proxies (Figure 6b and c) [7,24,58,85]. The R AK values of these two sites from 7 to 6 Ma changed from significant negative values before 7 Ma to positive or near-zero values after 6.5 Ma, which is the signal of increasing Taiwan sediment input. The R AK values raised before 6.5 Ma, because the erosion and sediment transport processes like gravity flow and contour current could exist even before Taiwan uplifted above the sea surface [1,89]. In addition, the 87Sr/86Sr values increased after 6.5 Ma, also indicating an increase in the relative input from Taiwan rivers (Figure 6b) [58]. Considering the timing of Taiwan uplift at ∼6.5 Ma, we suggest that the transport of Taiwan sediment to the northern SCS probably started around 6.5 Ma. At ∼6.5 Ma, the Taiwan sediments may contain a lot of coarse-grained quartz, feldspar, and mica instead of finer clay minerals due to the insufficient weathering, so the emergence of Taiwan would not influence the clay mineral assemblages drastically during the Late Miocene (Figure 6d).
![Figure 6
Temporal evolution of proxies linked to provenance change since the Late Miocene. (a) The R
AK values of IODP Site U1501 and ODP Sites 1144, 1146, and 1148. (b) 87Sr/86Sr values at ODP Site 1148 [7]. (c) ε
Nd values at ODP Site 1148 [58,85]. (d) Illite crystallinity values of ODP Site 1146 [24]. Blue and cyan horizontal bars represent ice volume in each hemisphere [86]. The timing of NHG formation is referred to in ref. [87], the timing of MPT is referred to in ref. [88]. The red line highlights the start of Taiwan orogen [9].](/document/doi/10.1515/geo-2022-0454/asset/graphic/j_geo-2022-0454_fig_006.jpg)
Temporal evolution of proxies linked to provenance change since the Late Miocene. (a) The R AK values of IODP Site U1501 and ODP Sites 1144, 1146, and 1148. (b) 87Sr/86Sr values at ODP Site 1148 [7]. (c) ε Nd values at ODP Site 1148 [58,85]. (d) Illite crystallinity values of ODP Site 1146 [24]. Blue and cyan horizontal bars represent ice volume in each hemisphere [86]. The timing of NHG formation is referred to in ref. [87], the timing of MPT is referred to in ref. [88]. The red line highlights the start of Taiwan orogen [9].
5.3 Provenance changes since the Pliocene
After the onset of Taiwan input at ∼6.5 Ma, the change of material source since the Pliocene was mainly controlled by sea level and precipitation instead of tectonic activities. Based on the R AK values (Figure 6a), we find that the most obvious provenance change is the long-term increase of the Pearl River input after 3.6 Ma. During the Mid-Pleistocene Transition (MPT) which occurred from 1.25 to 0.7 Ma [88], the Taiwan input relatively increased (Figure 6a). In addition, the fluctuation around 5 Ma could be related to the uplift of Dongsha, which mainly consists of basic igneous rocks, is very different from the Pearl River and Taiwan and similar to Luzon [90–92]. The weathering of Dongsha bedrocks could cause the rapid increase of smectite percentages of ODP Site 1146 from 5 to 3.6 Ma [6,93].
The increase of the Pearl River input after 3.6 Ma was mainly related to the growth of ice sheet. The Northern Hemisphere Glaciation (NHG) initiated at 3.6–3.3 Ma and greatly increased at 2.7 Ma [87]. Since 3.6 Ma, the lowered R AK of ODP Sites 1146 and 1148 and the increased illite crystallinity values of ODP Site 1146 both indicate an obvious increase of the Pearl River input (Figure 6a and d). We suggest that the lower sea level is the primary cause for increasing input from the Pearl River since the Late Pliocene. The sea level dropped about 50 m in the SCS after 3.6 Ma (Figure 4b) [53,54]. The water depth of the shelf edge is approximately 200 m [39], given the over 200 km wide continental shelf (Figures 1 and 7), the Pearl River mouth would get distinctly closer to the continental slope when sea level fell ∼100 m during glacials, and more sediments from the Pearl River would be delivered to the northern SCS slope [36,94]. In contrast, the Taiwan input is insensitive to sea level change because the shelf is much narrower (Figures 1 and 7) [1].
![Figure 7
The evolution of the source-to-sink system since the late Miocene. (a) Late Miocene (11 Ma); (b) Late Miocene (6 Ma); (c) Last Glacial Maximum (20 ka); (d) modern (0 ka). The Miocene paleogeographic maps are modified from ref. [84].](/document/doi/10.1515/geo-2022-0454/asset/graphic/j_geo-2022-0454_fig_007.jpg)
The evolution of the source-to-sink system since the late Miocene. (a) Late Miocene (11 Ma); (b) Late Miocene (6 Ma); (c) Last Glacial Maximum (20 ka); (d) modern (0 ka). The Miocene paleogeographic maps are modified from ref. [84].
During the MPT, the Taiwan input increased based on the positive R AK values of ODP Sites 1144, 1146, and 1148 (Figure 6a), we suggest the principal mechanisms are the NPIW expansion [42,95] and stronger seasonality according to the fluctuating smectite/kaolinite values after 1.25 Ma (Figure 4e) [38]. NPIW is shallow intermediate water that occupies a depth range of 300–800 m today [41], but its lower boundary deepened to >2 km around 0.9 Ma [95]. The enhancement of NPIW would improve the transport efficiency of Taiwan sediments as it flows westward along the northern SCS slope [18]. The expansion of NPIW ceased after the Mid-Brunhes Event at ∼0.43 Ma [96], contributing to the end of the high Taiwan input thereafter.
During the MPT, not only the NPIW but also the Mediterranean Outflow Water [97], the Indonesian Throughflow [98], and the North Atlantic Current [99] were strengthened, which were strongly related to the oceanographic and eustatic changes that occurred during the MPT [100]. After the MPT (<0.8 Ma), the sediment supply of the Pearl River enhanced again due to the high-amplitude (∼100 m) fluctuations of the global sea level (Figure 7). Furthermore, the sediment supply also increased in continental margins all over the world after the MPT, including the Mediterranean [101,102], the Southern Ocean [103], the Bohai Sea [104], and the Southwest Shelf of Australia [105].
In summary, the provenance changes of the northern SCS since the Late Miocene can be roughly divided into six stages: before 6.5 Ma, the sediment was mainly delivered from Pearl River (Figure 7a); after 6.5 Ma, due to the uplift of Taiwan, the poor-weathered Taiwan sediment began to enter the northern SCS (Figure 7b); from 6.5 to 3.6 Ma, the chemical weathering intensity in Taiwan gradually increased and started to change cooperatively with climate (Figure 4f); after 3.6 Ma, the sea level fell dramatically (Figure 4b) due to the NHG formation, and the Pearl River input rapidly increased; during the MPT, the Taiwan input increased because of the enhanced NPIW; and after the MPT, the glacial ages became longer and colder, the sea level reached the lowest stage (Figure 4b), the Pearl River input increased again (Figure 7c and d). In general, the chemical composition of sediments in the northern SCS was not controlled by climate during the Late Miocene when the sediment composition of Taiwan was still unstable. After the Pliocene, the coordinated variation between Taiwan sediment and the climate became more stabilized, so the climate change regained control over the relative input of each sediment source.
6 Conclusions
Provenance change has a profound effect on sediment composition in the northern SCS. By using the major elemental provenance proxies R AK and R KCN, combining with the mineral and isotope data, we reconstruct the evolutionary history of material sources on the northern SCS slope since the Late Miocene. We find that the northern SCS continental slope received sediment from Taiwan immediately after the Taiwan uplift at ∼6.5 Ma. The low weathering degree of the newly exposed Taiwan debris seems to result in a rapid decrease of the weathering intensity of the sediment on the northern SCS slope from 6.5 to 3.6 Ma. As the weathering condition in Taiwan became more stable around the Late Pliocene, the climate became a more crucial factor of provenance change. The relative input of Taiwan and the Pearl River is mainly controlled by sea level and precipitation after the late Pliocene. The expansion of Northern Hemisphere ice sheet around 3.6 Ma resulted in repeated falls in sea level and the increase of the relative input of Pearl River. During the MPT, the relative input of Taiwan increased due to the NPIW expansion. After the MPT, the ice sheet developed remarkably, the sea level reached an extremely low stand which can significantly strengthen the Pearl River input during the subsequent cold glacial periods.
Acknowledgements
We are grateful to the captains, scientists, officers, and crew of JOIDES Resolution for their cooperation in collecting sediment cores during the ODP Leg 184 in 1999 and IODP Expedition 367/368 in 2017. Jianghui Du (Eidgenössische Technische Hochschule Zürich) is appreciated for their constructive comments that help to improve this article.
-
Funding information: This work was jointly supported by National Natural Science Foundation of China (No. 41376043) and Guangdong Basic and Applied Basic Research Foundation (No. 2019A1515110896).
-
Author contributions: Z.H. and B.Q.H. conceived the idea of the study. Z.H. and B.Q.H. selected the samples. L.G. did analytical testing. Z.H. interpreted the results and wrote the article. L.G. and N.W. revised the manuscript and improved the language. All authors read and approved the final manuscript. The authors applied the SDC approach for the sequence of authors.
-
Conflict of interest: The authors state no conflict of interest in this article.
-
Data availability statement: The data used in this paper are open-source data available at https://doi.pangaea.de/10.1594/PANGAEA.937324.
References
[1] Liu Z, Zhao Y, Colin C, Stattegger K, Wiesner MG, Huh CA, et al. Source-to-sink transport processes of fluvial sediments in the South China Sea. Earth Sci Rev. 2016;153:238–73.10.1016/j.earscirev.2015.08.005Search in Google Scholar
[2] Sømme TO, Helland‐Hansen W, Martinsen OJ, Thurmond JB. Relationships between morphological and sedimentological parameters in source‐to‐sink systems: a basis for predicting semi‐quantitative characteristics in subsurface systems. Basin Res. 2009;21(4):361–87.10.1111/j.1365-2117.2009.00397.xSearch in Google Scholar
[3] Amorosi A, Sammartino I, Dinelli E, Campo B, Guercia T, Trincardi F, et al. Provenance and sediment dispersal in the Po-Adriatic source-to-sink system unraveled by bulk-sediment geochemistry and its linkage to catchment geology. Earth Sci Rev. 2022;234:104202.10.1016/j.earscirev.2022.104202Search in Google Scholar
[4] Milliman JD, Farnsworth KL. River Discharge to the Coastal Ocean: A Global Synthesis. Cambridge: Cambridge University Press; 2011.10.1017/CBO9780511781247Search in Google Scholar
[5] Wang P, Li Q, editors. The South China Sea: Paleoceanography and Sedimentology. Berlin: Springer Science & Business Media; 2009.10.1007/978-1-4020-9745-4Search in Google Scholar
[6] Wan S, Clift PD, Li A, Li T, Yin X. Geochemical records in the South China Sea: implications for East Asian summer monsoon evolution over the last 20 Ma. Geol Soc Spec Publ. 2010a;342(1):245–63.10.1144/SP342.14Search in Google Scholar
[7] Clift PD, Wan S, Blusztajn J. Reconstructing chemical weathering, physical erosion and monsoon intensity since 25 Ma in the northern South China Sea: a review of competing proxies. Earth Sci Rev. 2014;130:86–102.10.1016/j.earscirev.2014.01.002Search in Google Scholar
[8] Hu D, Clift PD, Wan S, Böning P, Hannigan R, Hillier S, et al. Testing chemical weathering proxies in Miocene – recent fluvial-derived sediments in the South China Sea. Geol Soc Spec Publ. 2016;429(1):45–72.10.1144/SP429.5Search in Google Scholar
[9] Chen WH, Huang CY, Lin YJ, Zhao Q, Yan Y, Chen D, et al. Depleted deep South China Sea δ13C paleoceanographic events in response to tectonic evolution in Taiwan–Luzon Strait since Middle Miocene. Deep Sea Res Part II Top Stud Oceanogr. 2015;122:195–225.10.1016/j.dsr2.2015.02.005Search in Google Scholar
[10] Chen WS, Yeh JJ, Syu SJ. Late Cenozoic exhumation and erosion of the Taiwan orogenic belt: new insights from petrographic analysis of foreland basin sediments and thermochronological dating on the metamorphic orogenic wedge. Tectonophysics. 2019;750:56–69.10.1016/j.tecto.2018.09.003Search in Google Scholar
[11] Malavieille J, Dominguez S, Lu CY, Chen CT, Konstantinovskaya E. Deformation partitioning in mountain belts: insights from analogue modelling experiments and the Taiwan collisional orogen. Geol Mag. 2021;158(1):84–103.10.1017/S0016756819000645Search in Google Scholar
[12] Lin AT, Watts AB, Hesselbo SP. Cenozoic stratigraphy and subsidence history of the South China Sea margin in the Taiwan region. Basin Res. 2003;15(4):453–78.10.1046/j.1365-2117.2003.00215.xSearch in Google Scholar
[13] Ding W, Li J, Han XQ, Li MB. Geomorphology, grain-size characteristics, matter source and forming mechanism of sediment waves on the ocean bottom of the northeast South China Sea. Acta Oceanol Sin. 2010;32(2):96–105 (in Chinese with English abstract).Search in Google Scholar
[14] Huang CY, Chen WH, Wang MH, Lin CT, Yang S, Li X, et al. Juxtaposed sequence stratigraphy, temporal-spatial variations of sedimentation and development of modern-forming forearc Lichi Mélange in North Luzon Trough forearc basin onshore and offshore eastern Taiwan: an overview. Earth Sci Rev. 2018;182:102–40.10.1016/j.earscirev.2018.01.015Search in Google Scholar
[15] Tsai CH, Shyu JB, Chung SL, Lee HY. Miocene sedimentary provenance and paleogeography of the Hengchun Peninsula, southern Taiwan: implications for tectonic development of the Taiwan orogen. J Asian Earth Sci. 2020;194:104032.10.1016/j.jseaes.2019.104032Search in Google Scholar
[16] Hall R. Cenozoic reconstructions of SE Asia and the SW Pacific: changing patterns of land and sea. Faunal Flor Migr Evol SE Asia-Australasia. 2001;1:35–56.Search in Google Scholar
[17] Gong C, Wang Y, Zhu W, Li W, Xu Q. Upper Miocene to Quaternary unidirectionally migrating deep-water channels in the Pearl River Mouth Basin, northern South China Sea unidirectionally migrating deep-water channels. AAPG Bull. 2013 Feb 1;97(2):285–308.10.1306/07121211159Search in Google Scholar
[18] You Y, Chern CS, Yang Y, Liu CT, Liu KK, Pai SC. The South China Sea, a cul-de-sac of North Pacific intermediate water. J Oceanogr. 2005;61(3):509–27.10.1007/s10872-005-0059-6Search in Google Scholar
[19] Nan F, Xue H, Yu F. Kuroshio intrusion into the South China Sea: a review. Prog Oceanogr. 2015;137:314–33.10.1016/j.pocean.2014.05.012Search in Google Scholar
[20] Wang X, Cai F, Sun Z, Li Q, Li A, Sun Y, et al. Late Miocene−Quaternary seismic stratigraphic responses to tectonic and climatic changes at the northeastern margin of the South China Sea. GSA Bull. 2022;134(9–10):2611–32.10.1130/B36224.1Search in Google Scholar
[21] Liu C, Clift PD, Murray RW, Blusztajn J, Ireland T, Wan S, et al. Geochemical evidence for initiation of the modern Mekong delta in the southwestern South China Sea after 8 Ma. Chem Geol. 2017a;451:38–54.10.1016/j.chemgeo.2017.01.008Search in Google Scholar
[22] Hu Z, Huang BQ, Liu LJ, Wang N. Spatiotemporal patterns of sediment deposition on the northern slope of the South China Sea in the last 150,000 years. J Palaeogeogr. 2021;10(1):1–7.10.1186/s42501-021-00102-3Search in Google Scholar
[23] Wei G, Liu Y, Li X, Shao L, Liang X. Climatic impact on Al, K, Sc and Ti in marine sediments: evidence from ODP Site 1144, South China Sea. Geochem J. 2003;37(5):593–602.10.2343/geochemj.37.593Search in Google Scholar
[24] Wan S, Li A, Clift PD, Wu S, Xu K, Li T. Increased contribution of terrigenous supply from Taiwan to the northern South China Sea since 3 Ma. Mar Geol. 2010b;278(1–4):115–21.10.1016/j.margeo.2010.09.008Search in Google Scholar
[25] Nasbitt HW, Young GM. Formation and diagenesis of weathering profiles. J Geol. 1989;97:129–47.10.1086/629290Search in Google Scholar
[26] McLennan SM. Weathering and global denudation. J Geol. 1993;101(2):295–303.10.1086/648222Search in Google Scholar
[27] von Eynatten H, Barceló-Vidal C, Pawlowsky-Glahn V. Modelling compositional change: the example of chemical weathering of granitoid rocks. Math Geol. 2003;35(3):231–51.10.1023/A:1023835513705Search in Google Scholar
[28] Boulay S, Colin C, Trentesaux A, Clain S, Liu Z, Lauer-Leredde C. Sedimentary responses to the Pleistocene climatic variations recorded in the South China Sea. Quat Res. 2007;68(1):162–72.10.1016/j.yqres.2007.03.004Search in Google Scholar
[29] Luan X, Zhang L, Peng X. Dongsha erosive channel on northern South China Sea shelf and its induced Kuroshio South China Sea branch. Sci China Earth Sci. 2012;55(1):149–58.10.1007/s11430-011-4322-ySearch in Google Scholar
[30] Wang D, Wang Q, Zhou W, Cai S, Li L, Hong B. An analysis of the current deflection around Dongsha Islands in the northern South China Sea. J Geophys Res Ocean. 2013;118(1):490–501.10.1029/2012JC008429Search in Google Scholar
[31] Shao L, Li X, Geng J, Pang X, Lei Y, Qiao P, et al. Deep water bottom current deposition in the northern South China Sea. Sci China Ser D Earth Sci. 2007;50(7):1060–6.10.1007/s11430-007-0015-ySearch in Google Scholar
[32] Zhu M, Graham S, Pang X, McHargue T. Characteristics of migrating submarine canyons from the middle Miocene to present: implications for paleoceanographic circulation, northern South China Sea. Mar Pet Geol. 2010;27(1):307–19.10.1016/j.marpetgeo.2009.05.005Search in Google Scholar
[33] Webster PJ. The role of hydrological processes in ocean–atmosphere interactions. Rev Geophys. 1994;32(4):427–76.10.1029/94RG01873Search in Google Scholar
[34] Huang W, Wang P. Sediment mass and distribution in the South China Sea since the Oligocene. Sci China Earth Sci. 2006;49(11):1147–55.10.1007/s11430-006-2019-4Search in Google Scholar
[35] Liu Z, Tuo S, Colin C, Liu JT, Huang CY, Selvaraj K, et al. Detrital fine-grained sediment contribution from Taiwan to the northern South China Sea and its relation to regional ocean circulation. Mar Geol. 2008;255(3–4):149–55.10.1016/j.margeo.2008.08.003Search in Google Scholar
[36] Liu J, Steinke S, Vogt C, Mohtadi M, De Pol-Holz R, Hebbeln D. Temporal and spatial patterns of sediment deposition in the northern South China Sea over the last 50,000 years. Palaeogeogr Palaeoclimatol Palaeoecol. 2017b;465:212–24.10.1016/j.palaeo.2016.10.033Search in Google Scholar
[37] Shao L, Cui Y, Stattegger K, Zhu W, Qiao P, Zhao Z. Drainage control of Eocene to Miocene sedimentary records in the southeastern margin of Eurasian Plate. GSA Bull. 2019;131(3–4):461–78.10.1130/B32053.1Search in Google Scholar
[38] Wan S, Li A, Clift PD, Stuut JB. Development of the East Asian monsoon: mineralogical and sedimentologic records in the northern South China Sea since 20 Ma. Palaeogeogr Palaeoclimatol Palaeoecol. 2007;254(3–4):561–82.10.1016/j.palaeo.2007.07.009Search in Google Scholar
[39] Zhong Y, Chen Z, Li L, Liu J, Li G, Zheng X, et al. Bottom water hydrodynamic provinces and transport patterns of the northern South China Sea: evidence from grain size of the terrigenous sediments. Cont Shelf Res. 2017;140:11–26.10.1016/j.csr.2017.01.023Search in Google Scholar
[40] Shaw PT, Chao SY. Surface circulation in the South China sea. Deep Sea Res Part I Oceanogr Res Pap. 1994;41(11–12):1663–83.10.1016/0967-0637(94)90067-1Search in Google Scholar
[41] Talley LD. Freshwater transport estimates and the global overturning circulation: shallow, deep and throughflow components. Prog Oceanogr. 2008;78(4):257–303.10.1016/j.pocean.2008.05.001Search in Google Scholar
[42] Knudson KP, Ravelo AC. North Pacific intermediate water circulation enhanced by the closure of the Bering Strait. Paleoceanography. 2015;10:1287–304.10.1002/2015PA002840Search in Google Scholar
[43] Qu T, Girton JB, Whitehead JA. Deepwater overflow through Luzon strait. J Geophys Res Ocean. 2006;111:C01002.10.1029/2005JC003139Search in Google Scholar
[44] Zhao W, Zhou C, Tian J, Yang Q, Wang B, Xie L, et al. Deep water circulation in the Luzon Strait. J Geophys Res Ocean. 2014;119(2):790–804.10.1002/2013JC009587Search in Google Scholar
[45] Sun Q, Cartwright J, Wu S, Zhong G, Wang S, Zhang H. Submarine erosional troughs in the northern South China Sea: evidence for Early Miocene deepwater circulation and paleoceanographic change. Mar Pet Geol. 2016;77:75–91.10.1016/j.marpetgeo.2016.06.005Search in Google Scholar
[46] Wang X, Zhuo H, Wang Y, Mao P, He M, Chen W, et al. Controls of contour currents on intra-canyon mixed sedimentary processes: insights from the Pearl River Canyon, northern South China Sea. Mar Geol. 2018;406:193–213.10.1016/j.margeo.2018.09.016Search in Google Scholar
[47] Chen H, Stow DA, Xie X, Ren J, Mao K, Gao Y, et al. Depositional architecture and evolution of basin-floor fan systems since the Late Miocene in the Northwest Sub-Basin, South China Sea. Mar Pet Geol. 2021;126:104803.10.1016/j.marpetgeo.2020.104803Search in Google Scholar
[48] Jian Z, Larsen HC, Alvarez Zarikian CA. The Expedition 368 Scientists. Expedition 368 Preliminary Report: South China Sea Rifted Margin. International Ocean Discovery Program; 2018.10.14379/iodp.pr.368.2018Search in Google Scholar
[49] Ma R, Liu C, Li Q, Jin X. Calcareous nannofossil changes in response to the spreading of the South China Sea basin during Eocene–Oligocene. J Asian Earth Sci. 2019;184:103963.10.1016/j.jseaes.2019.103963Search in Google Scholar
[50] Wei G, Li XH, Liu Y, Shao L, Liang X. Geochemical record of chemical weathering and monsoon climate change since the early Miocene in the South China Sea. Paleoceanography. 2006;21(4):PA4214.10.1029/2006PA001300Search in Google Scholar
[51] Gonneea ME, Paytan A. Phase associations of barium in marine sediments. Mar Chem. 2006;100(1–2):124–35.10.1016/j.marchem.2005.12.003Search in Google Scholar
[52] Zhao D, Wan S, Clift PD, Tada R, Huang J, Yin X, et al. Provenance, sea-level and monsoon climate controls on silicate weathering of Yellow River sediment in the northern Okinawa Trough during late last glaciation. Palaeogeogr Palaeoclimatol Palaeoecol. 2018;490:227–39.10.1016/j.palaeo.2017.11.002Search in Google Scholar
[53] Miller KG, Browning JV, Schmelz WJ, Kopp RE, Mountain GS, Wright JD. Cenozoic sea-level and cryospheric evolution from deep-sea geochemical and continental margin records. Sci Adv. 2020;6(20):eaaz1346.10.1126/sciadv.aaz1346Search in Google Scholar PubMed PubMed Central
[54] Pang X, Chen CM, Shi HS, Shu Y, Shao L, He M, et al. Response between relative sea-level change and the Pearl River deep-water fan system in the South China Sea. Earth Sci Front. 2005;12(3):167–77 (in Chinese with English abstract).Search in Google Scholar
[55] Larsen HC, Sun Z, Stock JM. The Expedition 367/368 Scientists. South China sea rifted margin. Proceedings of the International Ocean Discovery Program. 367/368: College Station, TX: (International Ocean Discovery Program); 2018.Search in Google Scholar
[56] Jiang HY, Li AC, Wan SM. Terrigenous mineralogical component in the sedments from SCS since 30 Ma and their geological significance. Stud. Mar. 2006;047(001):83–94 (in Chinese with English abstract).Search in Google Scholar
[57] Li AC, Huang J, Jiang HY, Wan SM. Sedimentary evolution in the northern slope of South China Sea since Oligocene and its responses to tectonics. Chin J Geophys. 2011;54(12):3233–45 (in Chinese with English abstract).10.1002/cjg2.1686Search in Google Scholar
[58] Clift P, Lee JI, Clark MK, Blusztajn J. Erosional response of South China to arc rifting and monsoonal strengthening; a record from the South China Sea. Mar Geol. 2002;184(3–4):207–26.10.1016/S0025-3227(01)00301-2Search in Google Scholar
[59] Clemens SC, Prell WL, Sun Y, Liu Z, Chen G. Southern Hemisphere forcing of Pliocene δ18O and the evolution of Indo‐Asian monsoons. Paleoceanography. 2008;23(4):PA4210.10.1029/2008PA001638Search in Google Scholar
[60] Steinke S, Groeneveld J, Johnstone H, Rendle-Bühring R. East Asian summer monsoon weakening after 7.5 Ma: evidence from combined planktonic foraminifera Mg/Ca and δ18O (ODP Site 1146; northern South China Sea). Palaeogeogr Palaeoclimatol Palaeoecol. 2010;289(1–4):33–43.10.1016/j.palaeo.2010.02.007Search in Google Scholar
[61] Cheng X, Zhao Q, Wang J, Jian Z, Xia P, Huang B, et al. 5. Data report: stable isotopes from sites 1147 and 1148. Age (Ma). 2004;1(1):3.10.2973/odp.proc.sr.184.223.2004Search in Google Scholar
[62] Clift PD. Asian monsoon dynamics and sediment transport in SE Asia. J Asian Earth Sci. 2020;195:104352.10.1016/j.jseaes.2020.104352Search in Google Scholar
[63] Israeli Y, Emmanuel S. Impact of grain size and rock composition on simulated rock weathering. Earth Surf Dyn. 2018;6(2):319–27.10.5194/esurf-6-319-2018Search in Google Scholar
[64] Andrews JE, Brimblecombe P, Jickells TD, Liss PS, Reid B. An introduction to environmental chemistry. Oxford: Blackwell Pub; 2016.Search in Google Scholar
[65] Holbourn AE, Kuhnt W, Clemens SC, Kochhann KG, Jöhnck J, Lübbers J, et al. Late Miocene climate cooling and intensification of southeast Asian winter monsoon. Nat Commun. 2018;9(1):1–3.10.1038/s41467-018-03950-1Search in Google Scholar PubMed PubMed Central
[66] Zhang H, Nie J, Liu X, Pullen A, Li G, Peng W, et al. Spatially variable provenance of the Chinese Loess Plateau. Geology. 2021;49(10):1155–9.10.1130/G48867.1Search in Google Scholar
[67] Sun D, Su R, Bloemendal J, Lu H. Grain-size and accumulation rate records from Late Cenozoic aeolian sequences in northern China: implications for variations in the East Asian winter monsoon and westerly atmospheric circulation. Palaeogeogr Palaeoclimatol Palaeoecol. 2008;264(1–2):39–53.10.1016/j.palaeo.2008.03.011Search in Google Scholar
[68] Shi Z, Liu X, An Z, Yi B, Yang P, Mahowald N. Simulated variations of eolian dust from inner Asian deserts at the mid-Pliocene, last glacial maximum, and present day: contributions from the regional tectonic uplift and global climate change. Clim Dyn. 2011;37(11):2289–301.10.1007/s00382-011-1078-1Search in Google Scholar
[69] Zhang W, Chen J, Ji J, Li G. Evolving flux of Asian dust in the North Pacific Ocean since the late Oligocene. Aeolian Res. 2016;23:11–20.10.1016/j.aeolia.2016.09.004Search in Google Scholar
[70] Du S, Xiang R, Liu J, Liu JP, Islam GA, Chen M. The present-day atmospheric dust deposition process in the South China Sea. Atmos Environ. 2020;223:117261.10.1016/j.atmosenv.2020.117261Search in Google Scholar
[71] Han W, Fang X, Berger A, Yin Q. An astronomically tuned 8.1 Ma eolian record from the Chinese Loess Plateau and its implication on the evolution of Asian monsoon. J Geophys Res Atmos. 2011;116:D24114.10.1029/2011JD016237Search in Google Scholar
[72] Gai C, Liu Q, Roberts AP, Chou Y, Zhao X, Jiang Z, et al. East Asian monsoon evolution since the late Miocene from the South China Sea. Earth Planet Sci Lett. 2020;530:115960.10.1016/j.epsl.2019.115960Search in Google Scholar
[73] Lu H, Wang X, Wang X, Chang X, Zhang H, Xu Z, et al. Formation and evolution of Gobi Desert in central and eastern Asia. Earth Sci Rev. 2019;194:251–63.10.1016/j.earscirev.2019.04.014Search in Google Scholar
[74] Rea DK, Hovan SA. Grain size distribution and depositional processes of the mineral component of abyssal sediments: lessons from the North Pacific. Paleoceanography. 1995;10(2):251–8.10.1029/94PA03355Search in Google Scholar
[75] Jiang F, Zhou Y, Nan Q, Zhou Y, Zheng X, Li T, et al. Contribution of Asian dust and volcanic material to the western Philippine Sea over the last 220 kyr as inferred from grain size and Sr–Nd isotopes. J Geophys Res Ocean. 2016;121(9):6911–28.10.1002/2016JC012000Search in Google Scholar
[76] Miri A, Dragovich D, Dong Z. Vegetation morphologic and aerodynamic characteristics reduce aeolian erosion. Sci Rep. 2017;7(1):1–9.10.1038/s41598-017-13084-xSearch in Google Scholar PubMed PubMed Central
[77] Liu J, Kimura R, Miyawaki M, Kinugasa T. Effects of plants with different shapes and coverage on the blown-sand flux and roughness length examined by wind tunnel experiments. Catena. 2021;197:104976.10.1016/j.catena.2020.104976Search in Google Scholar
[78] Sun X, Li X, Luo Y, Chen X. The vegetation and climate at the last glaciation on the emerged continental shelf of the South China Sea. Palaeogeogr Palaeoclimatol Palaeoecol. 2000;160(3–4):301–16.10.1016/S0031-0182(00)00078-XSearch in Google Scholar
[79] Wan S, Clift PD, Zhao D, Hovius N, Munhoven G, France-Lanord C, et al. Enhanced silicate weathering of tropical shelf sediments exposed during glacial lowstands: a sink for atmospheric CO2. Geochim Cosmochim Acta. 2017;200:123–44.10.1016/j.gca.2016.12.010Search in Google Scholar
[80] Clift PD, Webb AA. A history of the Asian monsoon and its interactions with solid Earth tectonics in Cenozoic South Asia. Geol Soc Spec Publ. 2019;483(1):631–52.10.1144/SP483.1Search in Google Scholar
[81] Tang H. The spatio-temporal evolution of the Asian monsoon climate in the Late Miocene and its causes: a regional climate model study. Helsinki: Department of Geosciences and Geography A21; 2013.Search in Google Scholar
[82] Miao Y, Warny S, Clift PD, Liu C, Gregory M. Evidence of continuous Asian summer monsoon weakening as a response to global cooling over the last 8 Ma. Gondwana Res. 2017;52:48–58.10.1016/j.gr.2017.09.003Search in Google Scholar
[83] Haq BU, Hardenbol JA, Vail PR. Chronology of fluctuating sea levels since the Triassic. Science. 1987;235(4793):1156–67.10.1126/science.235.4793.1156Search in Google Scholar PubMed
[84] Collins DS, Avdis A, Allison PA, Johnson HD, Hill J, Piggott MD, et al. Tidal dynamics and mangrove carbon sequestration during the Oligo–Miocene in the South China Sea. Nat Commun. 2017;8(1):1–2.10.1038/ncomms15698Search in Google Scholar PubMed PubMed Central
[85] Li XH, Wei G, Shao L, Liu Y, Liang X, Jian Z, et al. Geochemical and Nd isotopic variations in sediments of the South China Sea: a response to Cenozoic tectonism in SE Asia. Earth Planet Sci Lett. 2003;211(3–4):207–20.10.1016/S0012-821X(03)00229-2Search in Google Scholar
[86] Westerhold T, Marwan N, Drury AJ, Liebrand D, Agnini C, Anagnostou E, et al. An astronomically dated record of Earth’s climate and its predictability over the last 66 million years. Science. 2020;369(6509):1383–7.10.1126/science.aba6853Search in Google Scholar PubMed
[87] De Schepper S, Gibbard PL, Salzmann U, Ehlers J. A global synthesis of the marine and terrestrial evidence for glaciation during the Pliocene Epoch. Earth Sci Rev. 2014;135:83–102.10.1016/j.earscirev.2014.04.003Search in Google Scholar
[88] Hönisch B, Hemming NG, Archer D, Siddall M, McManus JF. Atmospheric carbon dioxide concentration across the mid-Pleistocene transition. Science. 2009;324(5934):1551–4.10.1126/science.1171477Search in Google Scholar PubMed
[89] Porz L, Zhang W, Hanebuth TJ, Schrum C. Physical processes controlling mud depocenter development on continental shelves–geological, oceanographic, and modeling concepts. Mar Geol. 2021;432:106402.10.1016/j.margeo.2020.106402Search in Google Scholar
[90] Lüdmann T, Wong HK, Wang P. Plio–Quaternary sedimentation processes and neotectonics of the northern continental margin of the South China Sea. Mar Geol. 2001;172(3–4):331–58.10.1016/S0025-3227(00)00129-8Search in Google Scholar
[91] Yan YI, Xia B, Lin GE, Carter A, Hu X, Cui X, et al. Geochemical and Nd isotope composition of detrital sediments on the north margin of the South China Sea: provenance and tectonic implications. Sedimentology. 2007;54(1):1–7.10.1111/j.1365-3091.2006.00816.xSearch in Google Scholar
[92] Luan X, Ran W, Wang K, Wei X, Shi Y, Zhang H. New interpretation for the main sediment source of the rapidly deposited sediment drifts on the northern slope of the South China Sea. J Asian Earth Sci. 2019;171:118–33.10.1016/j.jseaes.2018.11.004Search in Google Scholar
[93] Wang F, Ding W. How did sediments disperse and accumulate in the oceanic basin, South China Sea. Mar Pet Geol. 2023;147:105979.10.1016/j.marpetgeo.2022.105979Search in Google Scholar
[94] Huang J, Wan S, Xiong Z, Zhao D, Liu X, Li A, et al. Geochemical records of Taiwan-sourced sediments in the South China Sea linked to Holocene climate changes. Palaeogeogr Palaeoclimatol Palaeoecol. 2016;441:871–81.10.1016/j.palaeo.2015.10.036Search in Google Scholar
[95] Kender S, Ravelo AC, Worne S, Swann GE, Leng MJ, Asahi H, et al. Closure of the Bering Strait caused mid-Pleistocene transition cooling. Nat Commun. 2018;9(1):1–1.10.1038/s41467-018-07828-0Search in Google Scholar PubMed PubMed Central
[96] Jonas AS, Kars MA, Bauersachs T, Rübsam W, Schwark L. Decoupling of Nw Pacific from global climate evolution linked to the Mid-Pleistocene Transition and Mid-Brunhes Event. In 29th International Meeting on Organic Geochemistry; 2019. Vol. 2019. Issue 1. p. 1–2. European Association of Geoscientists & Engineers10.3997/2214-4609.201903064Search in Google Scholar
[97] Guo Q, Li B, Voelker AH, Kim JK. Mediterranean outflow water dynamics across the middle Pleistocene transition based on a 1.3 million-year benthic foraminiferal record off the Portuguese margin. Quat Sci Rev. 2020;247:106567.10.1016/j.quascirev.2020.106567Search in Google Scholar
[98] Auer G, Petrick B, Martinez-Garcia A, Mamo B, Reuning L, De Vleeschouwer D, et al. The evolution of the Leeuwin Current and its Undercurrent during the Middle Pleistocene Transition – insights from multiproxy productivity records. Geophys Res Abstr. 2019;21:1–1.Search in Google Scholar
[99] Barker S, Zhang X, Jonkers L, Lordsmith S, Conn S, Knorr G. Strengthening Atlantic inflow across the Mid‐Pleistocene Transition. Paleoceanogr Paleoclimatol. 2021;36(4):e2020PA004200.10.1029/2020PA004200Search in Google Scholar
[100] Chalk TB, Hain MP, Foster GL, Rohling EJ, Sexton PF, Badger MP, et al. Causes of ice age intensification across the Mid-Pleistocene Transition. PNAS. 2017;114(50):13114–9.10.1073/pnas.1702143114Search in Google Scholar PubMed PubMed Central
[101] Pellegrini C, Maselli V, Trincardi F. Pliocene–Quaternary contourite depositional system along the south-western Adriatic margin: changes in sedimentary stacking pattern and associated bottom currents. Geo Mar Lett. 2016;36(1):67–79.10.1007/s00367-015-0424-4Search in Google Scholar
[102] Gauchery T, Rovere M, Pellegrini C, Cattaneo A, Campiani E, Trincardi F. Factors controlling margin instability during the plio-quaternary in the Gela Basin (Strait of Sicily, Mediterranean Sea). Mar Pet Geol. 2021;123:104767.10.1016/j.marpetgeo.2020.104767Search in Google Scholar
[103] Hillenbrand CD, Kuhn G, Frederichs T. Record of a Mid-Pleistocene depositional anomaly in West Antarctic continental margin sediments: an indicator for ice-sheet collapse? Quat Sci Rev. 2009;28(13–14):1147–59.10.1016/j.quascirev.2008.12.010Search in Google Scholar
[104] Shi X, Yao Z, Liu Q, Larrasoaña JC, Bai Y, Liu Y, et al. Sedimentary architecture of the Bohai Sea China over the last 1 Ma and implications for sea-level changes. Earth Planet Sci Lett. 2016;451:10–21.10.1016/j.epsl.2016.07.002Search in Google Scholar
[105] Deik H, Reuning L, Courtillat M, Petrick B, Bassetti MA. The sedimentary record of Quaternary glacial to interglacial sea‐level change on a subtropical carbonate ramp: Southwest Shelf of Australia. Sedimentology. 2021;68(2):593–608.10.1111/sed.12793Search in Google Scholar
© 2022 the author(s), published by De Gruyter
This work is licensed under the Creative Commons Attribution 4.0 International License.