Next Article in Journal
Examining the Impact of Long-Term Industrialization on the Trace Metal Contaminants Distribution in Seawater of the Pula Bay, Croatia
Next Article in Special Issue
Mid-Deep Circulation in the Western South China Sea and the Impacts of the Central Depression Belt and Complex Topography
Previous Article in Journal
Point-to-Point-Based Optimization Method of Ballast Water Allocation for Revolving Floating Cranes with Experimental Verification
Previous Article in Special Issue
Review of Asymmetric Seafloor Spreading and Oceanic Ridge Jumps in the South China Sea
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Astronomical Time Scale of the Late Pleistocene in the Northern South China Sea Based on Carbonate Deposition Record

School of Earth Sciences, Key Laboratory of Mineral Resources in Western China (Gansu Province), Lanzhou University, Lanzhou 730000, China
*
Authors to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2024, 12(3), 438; https://0-doi-org.brum.beds.ac.uk/10.3390/jmse12030438
Submission received: 1 February 2024 / Revised: 21 February 2024 / Accepted: 24 February 2024 / Published: 1 March 2024

Abstract

:
Variations in solar insolation caused by changes in the Earth’s orbit—specifically its eccentricity, obliquity, and precession—can leave discernible marks on the geologic record. Astrochronology leverages these markers to establish a direct connection between chronological measurements and different facets of climate change as recorded in marine sediments. This approach offers a unique window into the Earth’s climate system and the construction of high-resolution, continuous time scales. Our study involves comprehensive bulk carbonate analyses of 390 discrete samples from core SCS1, which was retrieved from the deep-sea floor of the northern South China Sea. By utilizing carbonate stratigraphic data, we have developed a carbonate stratigraphic age model. This was achieved by aligning the carbonate sequence from core SCS1 with the established carbonate standard stratigraphic time scale of the South China Sea. Subsequently, we construct an astronomically tuned time scale based on this age model. Our findings indicate that sediment records in this core have been predominantly influenced by a 20,000-year cycle (precession cycle) throughout the Late Pleistocene. We have developed an astronomical time scale extending back approximately 110,000 years from the present, with a resolution of 280 years, by tuning the carbonate record to the precession curve. Time-domain spectral analysis of the tuned carbonate time series, alongside the consistent comparability of the early Holocene low-carbonate event (11–8 kyr), underscores the reliability of our astronomical time scale. Our age model exposes intricate variations in carbonate deposition, epitomizing a typical “Pacific-type” carbonate cycle. Previous research has illustrated that precession forcing predominantly influences productivity changes in the South China Sea. The pronounced precession-related cycle observed in our record suggests that changes in productivity significantly impact carbonate content in the area under study. Furthermore, the clear precession period identified in the carbonate record of core SCS1 reflects the response of low-latitude processes to orbital parameters, implying that carbonate deposition and preservation in core SCS1 are chiefly influenced by the interplay between the Intertropical Convergence Zone (ITCZ) and the monsoon system within the precession band. Our astronomical time scale is poised to enhance paleoceanographic, paleoclimatic, and correlation studies further. Additionally, the independent evidence we provide for using proxy records for astronomical age calibration of marine sediments lends additional support to similar methods of astronomical tuning.

1. Introduction

Deep-sea sediment sequences are crucial documents of past global climate changes, serving as primary carriers of high-resolution, long-term records of marine history during the Pliocene–Pleistocene epochs [1,2,3]. The South China Sea (SCS), in particular, stands out due to its high sedimentation rates and favorable conditions for carbonate preservation, making it a rich archive of both regional and global climate changes. This provides an optimal setting for conducting detailed paleoceanographic studies [4,5]. Research has extensively explored the environmental evolution within the SCS since the Quaternary, revealing significant insights. For instance, analyses of cores ODP 1147/48 and ODP 1146 have delineated variations in monsoonal activity, while core MD12-3429 has shed light on the long-term impacts of sea level changes and Kuroshio intrusion on terrestrial sediment transport in the northeastern SCS [6,7,8]. The Kuroshio is an important part of the western boundary current of the northern Pacific Ocean, with high temperature and salinity characteristics [9,10]. Additionally, the U 37 K SST record has provided valuable data on glacial and interglacial temperature fluctuations over the last 150,000 years [11]. Despite these advancements, dating sedimentary records and analyzing paleoenvironmental climates in the SCS remain challenging due to the complex depositional environment and varied material sources [12,13].
An independent, orbitally scaled chronological framework is pivotal for accurately identifying and understanding shifts in paleoclimate and paleoenvironment [4,14]. To date, several methods have been applied to establish the chronological framework of the SCS. Zhong et al. and Lin et al. used optically stimulated luminescence of quartz to establish the chronological framework of the ChaoShan Delta and Pearl River Delta, respectively [15,16]. However, in marine depositional environments, the depositional process is complex. The transportation distance, particle size, depositional environment, and other factors can make the degree of solar bleaching of quartz or feldspar difficult to evaluate [14,17]. AMS 14C dating of planktonic foraminifera (preferably single species) has also been widely used to construct chronological frameworks for sedimentary sequences in the SCS [18,19,20,21]. However, merely relying on 14C to construct accurate chronologies for the SCS records is problematic. For instance, radiocarbon dating dates back to ~50 kyr ago [22]. It cannot be applied to sediments beyond this dating range. In addition, different planktonic foraminiferal species from the same sediment depth may have significantly different radiocarbon ages due to the effects of sediment redeposition, recrystallization, and incorporation of secondary radiocarbon [23]. Studies have also shown that bioturbation, carbonate dissolution, or chemical erosion may significantly affect radiocarbon-based age scales [24,25,26,27]. Finally, converting marine radiocarbon dates to calendar ages requires knowledge of the surface reservoir age. However, the surface reservoir age variations remain poorly constrained [28]. Beyond radiocarbon dating, matching oxygen isotopes from foraminifera to the global oxygen isotope stack offers another strategy for dating, extending back millions of years [29]. However, this method can only be used in areas where foraminifera are a common component of sediments. Furthermore, a characteristic species’ first or last occurrence can constrain the sediment age [4]. However, it is imprecise because of the significant provincialism of organisms used for biostratigraphic datums [2]. Similarly, while absolute ages can sometimes be inferred from ash layers or other distinctive sediments, the practical application of these methods is often hindered by the absence of well-preserved carbonates or suitable microfossils in many oceanic regions [30]. Except as mentioned above, the paleomagnetic dating techniques provide an effective method for deep-sea sediments [31] and a chronological framework for paleoenvironmental studies in the SCS [4,32]. However, paleomagnetic stratigraphic age control points are relatively limited, and traditional paleomagnetic stratigraphic methods are not applicable to sediments younger than the Matuyama–Brunhes boundary (~780 ka) [33]. Even though relative paleointensity (RPI) methods are now becoming increasingly popular for younger sediments (<780 ka), the RPI model remains ambiguous due to variations in depositional rates without further constraints from other dating methods [34,35]. Next, changes in magnetic minerals in anoxic environments can also record chemical remanent magnetization, which can confound previous interpretations of paleomagnetic stratigraphy. Given the varying age ranges and limitations of these dating methods [14], employing multi-parameter approaches is often necessary to ensure the accuracy of marine sediment chronologies [8,36].
In recent years, astrochronology has emerged as a pivotal method for constructing age models across diverse climatic records, including eolian, abyssal sediments, and continental sediments [4,37,38,39,40,41]. The foundational theory of astronomical climate, proposed by Milutin Milankovitch in the early 20th century, has paved the way for the development of precise, high-resolution astronomical chronologies [42,43]. The landmark study by Hays et al. [44] highlighted the significant role of Earth’s orbital geometry in driving climate change, a phenomenon well-documented within sedimentary records. Subsequent research has consistently shown that Quaternary climate changes are closely aligned with variations in the Earth’s orbit, forming the basis for paleoclimatic age models through astronomical calibration [3,45]. Astrochronology involves identifying Milankovitch cycles within climate proxies from sediment records and aligning these orbital signals with astronomical parameter models to establish a continuous, high-resolution geological chronological framework [46]. This method offers potentially higher resolution and accuracy than traditional time scales based on linear interpolation between geomagnetic reversals or radiometrically dated points [47,48]. Most new age models have been created by correlating isotopic records with orbital target curves [2,47,49,50]. In cases where isotope data are unavailable, other climate indicators such as magnetization, natural gamma, color reflectance, mineral content, elemental content, or carbonate content have been successfully employed [2,4,47,51,52,53,54,55]. Notably, astronomically calibrated carbonate stratigraphy has proven to be an effective tool for constructing high-resolution age models for oceanic deposits [52,54,56,57].
Astrochronology has been successfully applied in the SCS, reconstructing astronomical time scales for deep-sea sedimentary sequences [2,4,47,51]. Astronomical tuning age models for the region have utilized various signal curves and target curves, depending on the core location. For instance, cores ODP 1148, IODP U1499, and U1505 were tuned using combinations of natural gamma radiation, color spectral reflectance, and/or magnetic susceptibility, with global oxygen isotope curves, eccentricity, obliquity, and/or precession serving as target curves [2,4,51]. Additionally, cores ODP 1145 and 1143 employed K/Si ratios, benthic foraminiferal oxygen isotope records, and hematite to goethite content ratio (Hm/Gt) as the tuning signals, with the Northern Hemisphere summer insolation record as the primary tuning target [1,47,55].
The suitability of carbonate stratigraphy and astrochronology for core SCS1 in the SCS is supported by several factors: (a) the widespread availability of carbonate content as a classical paleoenvironmental proxy [58,59,60,61,62]; (b) the established link between carbonate content variations and glacial–interglacial cycles [60,63]; (c) the influence of monsoon variability on carbonate deposition and preservation [55], and (d) the significant effects of Milankovitch cycle-induced climatic oscillations on glacial–interglacial cycles, monsoon variability, and ocean productivity [64,65,66,67,68]. The change in carbonate content is affected by dilution and productivity [55]. Both dilution and productivity are related to monsoon variability [55]. Precession forcing has long been considered to be the main force of monsoon variation on orbital time scales [66,68]. It redistributes or splits seasonal incoming solar radiation, especially at low latitudes where monsoon systems prevail [64,65,66,67]. The summer and winter monsoons both have an impact on carbonate content through productivity and dilution [55]. Thus, the cycle of low-latitude solar radiation is also transmitted to the oceanic carbonate deposition record through the monsoon as a link. Furthermore, the periodic changes in solar insolation with latitude and season caused by changes in the Earth’s orbital parameters are the fundamental driving force for climate change and glacial cycles on a millennium to ten-thousand-year scale [69]. The periodicity controlled by the precession of monsoon variability is considered to be influenced by the Pliocene/Pleistocene glacial–interglacial cycle [55]. At the same time, the carbonate content during the late Pleistocene is also related to the glacial–interglacial cycle [55,60]. These factors suggest that Earth’s orbital parameter changes directly impact the deposition and preservation of carbonates in sediments, reflected in the carbonate and other biogeochemical records of ocean sediments. At present, the core SCS1 in the SCS has been obtained; however, different dating methods have different age ranges and constraints, such as the traditional paleomagnetic methods, which can only be relied on to decipher stratigraphic chronology [3]. Thus, the effectiveness of the astronomical tuning method in improving the resolution of the chronostratigraphy of the SCS requires further study.
This study aims to establish a continuous Late Pleistocene astronomically tuned carbonate stratigraphy of core SCS1 from the northern SCS. We present a high-resolution carbonate record from core SCS1, establish an initial untuned carbonate stratigraphic age model, and construct an astronomical time scale by tuning the carbonate content series to Earth’s orbital parameters. Through comparison with low-carbonate events in other SCS cores, we aim to verify the accuracy of the constructed astronomical time scale and discuss the influence of precession forcing on carbonate content variation.

2. Regional Setting

The SCS, framed by China, the Indo-China Peninsula, and the Taiwan–Philippine Island Arc, is a significant geographical feature in the western Pacific (Figure 1). This semi-enclosed deep-sea basin, the largest marginal sea in the western Pacific, spans approximately 3.5 million square kilometers [47]. It boasts an average depth of around 1250 m, plunging to over 5000 m at its deepest points. The basin is characterized by expansive shelf areas in the northern and southwestern regions, deep-sea basins in the central and eastern parts, and narrow straits in the southwest and northeast, facilitating its connection to the East China Sea and the Okinawa Trough via the Taiwan Strait to the north and to the open Pacific Ocean through the Bashi Strait to the east [70].
The climate of the region is predominantly controlled by the East Asian monsoon system, leading to distinct seasonal weather patterns across the SCS (Figure 1). Winters are marked by cold, dry conditions with prevailing northeast winds, while summers bring warm, wet weather with southwest winds, attributed to the seasonal reversal of wind patterns due to the Asian winter and summer monsoon circulations [47]. During winter, as northern Asia cools and high-pressure systems develop, northeast monsoon winds sweep across the sea. Conversely, the arrival of summer sees central China warming, low pressure systems forming, and the onset of gentle southwesterly winds [13]. The intense winter monsoon persists for about six months, from November to April, while the relatively milder summer monsoon spans approximately four months, from mid-May to mid-September [71].
The northern part of the SCS is influenced by a series of surface circulation and the joint influence of the Kuroshio SCS branch and the West Pacific Deep Current (Figure 1). Monsoons play a significant role in the general surface circulation of the SCS. The surface circulation is also strongly influenced by the Kuroshio Current intrusion in the northern part [72]. During the winter, the Kuroshio Current takes a direct route from northeast Luzon toward southwest Taiwan, continuing westward along the northern slope of the SCS. In summer, the current shifts southward, its flow markedly weakening [73]. Additionally, the deep-water current enters the SCS through the Luzon Strait, deflecting northwest before turning southwest along the continental margin of southeastern China, thus forming the SCS Contour Current [74].
Figure 1. Map showing the position of core SCS1, marked by a red star, within the context of the monsoon winds and current systems prevalent in the northern SCS. The depiction of monsoon winds is based on Webster (1994) [75], while the mapping of major surface currents follows the work of Fang et al. (1998) [72]. The surface currents in the Gulf of Tonkin are derived from studies by Xia et al. (2001) and Wu et al. (2008) [76,77], with the Kuroshio Current information adapted from Caruso et al. (2006) [78]. Deep-water currents are inferred from the combined research of Qu et al. (2006), Wang et al. (2011), and Zhao et al. (2014) [74,79,80]. The currents are numbered 1 to 9, indicating winter (depicted in black) and summer (depicted in red) surface currents, which include: (1) Loop Current, (2) SCS Branch of the Kuroshio, (3) NW Luzon Cyclonic Gyre, (4) NW Luzon Cyclonic Eddy, (5) NW Luzon Coastal Current, (6) SCS Warm Current, (7) Guangdong Coastal Current, (8) SE Vietnam Offshore Current, and (9) Gulf of Tonkin Surface Current. The numbers 10 to 12 represent deep-water currents: (10) Luzon Deep Current, (11) SCS Contour Current, and (12) Deep Cyclonic Current. The base map is constructed using ETOPO1 data.
Figure 1. Map showing the position of core SCS1, marked by a red star, within the context of the monsoon winds and current systems prevalent in the northern SCS. The depiction of monsoon winds is based on Webster (1994) [75], while the mapping of major surface currents follows the work of Fang et al. (1998) [72]. The surface currents in the Gulf of Tonkin are derived from studies by Xia et al. (2001) and Wu et al. (2008) [76,77], with the Kuroshio Current information adapted from Caruso et al. (2006) [78]. Deep-water currents are inferred from the combined research of Qu et al. (2006), Wang et al. (2011), and Zhao et al. (2014) [74,79,80]. The currents are numbered 1 to 9, indicating winter (depicted in black) and summer (depicted in red) surface currents, which include: (1) Loop Current, (2) SCS Branch of the Kuroshio, (3) NW Luzon Cyclonic Gyre, (4) NW Luzon Cyclonic Eddy, (5) NW Luzon Coastal Current, (6) SCS Warm Current, (7) Guangdong Coastal Current, (8) SE Vietnam Offshore Current, and (9) Gulf of Tonkin Surface Current. The numbers 10 to 12 represent deep-water currents: (10) Luzon Deep Current, (11) SCS Contour Current, and (12) Deep Cyclonic Current. The base map is constructed using ETOPO1 data.
Jmse 12 00438 g001

3. Materials and Methods

3.1. Core Description

Core SCS1, measuring 7.82 m in length, was collected from a deep-sea basin in the northern SCS using a gravity column sampler on 19 June 2017, at coordinates 18°30′22.56″ N, 116°16′5.28″ E and a water depth of 3770 m (Figure 1). The core is lithologically consistent, comprising gray-green clay silt (Figure 2). Its sedimentary profile is marked by continuous sedimentation, devoid of any discernible hiatuses, turbidite layers, or signs of biological disturbance.

3.2. Analysis of Carbonate (%) Content

In the sediments of the SCS, carbonate, primarily derived from calcareous microfossils such as foraminifera and nannoplankton, as well as from coral reefs, is the predominant biogenic material [70]. Core SCS1 was sampled at 2 cm intervals, yielding 390 samples for carbonate content analysis. The prevalent method for determining the carbonate content in sediment involves treating samples with hydrochloric acid, followed by the measurement of carbon dioxide release. This volume of carbon dioxide is then quantified as the carbonate concentration (%), employing the Scheibler method, with an accuracy of within ±1% [81].

3.3. Time Series Analysis

The time series analysis focused on the higher-resolution carbonate data series. The initial step involved using linear interpolation to resample the data, ensuring even spacing. To accentuate astronomical signals by mitigating low-frequency, long-term trends, the carbonate series underwent pre-whitening, subtracting a 35% weighted average through the LOWESS method [82]. The method emphasizes short-term changes to resolve high-frequency parameters from the strong red noise background signal common in stratigraphic data. Subsequently, the presence of cycles in the raw data was assessed using the 2π multi-taper method (MTM) for power spectral analysis [83], with spectral estimations compared against robust red noise models [84]. Evolutionary spectral analysis was conducted using a sliding window fast Fourier transform (FFT) spectrogram, enabling dynamic tracking of frequency shifts in the carbonate series over time [85]. Gaussian bandpass filtering isolated the interpreted Milankovitch cycles from the carbonate series. These procedures were executed using Acycle v2.4.1 software [86], with further details and guidance available at www.mingsongli.com/acycle (accessed on 6 March 2023). The final step involved tuning the carbonate series’ ~20 kyr component to align with the precession astronomical solution, utilizing QAnaly Series v1.5.0 software for the correlation between the carbonate record and the astronomical solution [87,88].

4. Results and Analyses

4.1. Carbonate Stratigraphy Comparison

The carbonate content of core SCS1 is generally low, ranging from 2.8% to 38.0%, with an average value of 13.8%. According to the level and change trend of carbonate content, this variability is categorized into four distinct stages (Figure 2): Stage I (7.82–5.2 mbsf): the carbonate content fluctuates between 10.3% and 15.1%, averaging 12.7%. This stage is marked by relatively low but gradually increasing carbonate levels from 6.58 to 6.02 mbsf, exhibiting minimal variability. Stage II (5.20–3.10 mbsf): here, the carbonate content spans 10.4% to 21.1%, with an average of 15.3%. This stage shows moderate variability and an overall increase in carbonate content compared to Stage. Stage III (3.10–0.98 mbsf): carbonate levels oscillate between 9.5% and 38.0%, with an average of 14.5%. This stage features significant variability, with lower content observed from 3.10 to 1.66 mbsf and higher levels noted between 1.66 to 0.98 mbsf. Stage IV (0.98–0.02 mbsf): this stage sees the carbonate content ranging from 2.8% to 20.1%, averaging 11.8%. Characterized by large fluctuations, the carbonate content generally decreases towards shallower depths, with the lowest value recorded in Stage IV. Periodic fluctuations across various scales in the data support the feasibility of cyclostratigraphy analysis.
Carbonate content is crucial for understanding deep-water sediments and offers valuable insights into marine paleoenvironmental and paleoclimate changes. It also reflects the historical interplay between climate change and land–sea interactions through the relative content of terrestrial detritus and marine authigenic carbonate deposits [89]. Dividing sedimentary strata based on carbonate content offers an effective alternative when foraminiferal oxygen isotope values are absent, a technique extensively applied in the Pacific since the Quaternary [58,59,61,90]. However, the standard carbonate stratigraphic time scale in the Pacific differs from that in the SCS. Given the ventilation of the deep SCS by the same Pacific Ocean deep-water mass through the Luzon Strait, any changes in seawater chemistry are recorded in the basin’s carbonate system [58,62,74]. Consequently, a standard carbonate stratigraphic time scale for the SCS was established by comparing extensive carbonate data with the SO49-37KL core, introducing the naming convention of the high-resolution standard carbonate stratigraphic event as SB events for the SCS, with numbers indicating specific events [58]. It is widely comparable in the SCS [58,90,91].
In this study, we compared the carbonate content of core SCS1 with the established SCS carbonate stratigraphic time scale to identify main-cycle events and establish an initial chronological framework. At a depth of 5.14 mbsf in core SCS1 (Figure 2), AMS 14C dating of the carbonate fraction in the sediment was conducted, yielding an age of 42,950 ± 610 years. After calibration using the Calib 8.20 program and Marine 20 dataset [92], with a regional reservoir correction, the calibrated age is 42,733 years. Due to some limitations of the AMS 14C dating of marine sediments mentioned earlier, this dating result serves as a reference for the carbonate stratigraphic comparison. The identified carbonate stratigraphic events—SB2.0, SB3.0, and SB5.0—correspond to ages of approximately 12 kyr, 24 kyr, and 74 kyr, respectively, corresponding to core depths of 98 cm, 310 cm, and 614 cm, in that order (Table 1; Figure 3 and Figure 4). These events align with the known stratigraphic divisions of the adjacent SO50-29KL core, demonstrating a high similarity in fluctuation trends (Table 1; Figure 3 and Figure 4). Further comparison with twenty high-resolution cores from the SCS confirms the reliability of the carbonate stratigraphic division results for core SCS1 (Table 1; Figure 3 and Figure 4). This comparison reveals consistent carbonate content changes across the region, with distinct patterns observed near identified events. However, regional sedimentation and sampling resolution may account for slight variations in the curve changes.
Table 1. Sediment cores from the SCS that were used in this study.
Table 1. Sediment cores from the SCS that were used in this study.
CoreLatitude (N)Longitude (E)Water Depth (m)References
SO49-37KL17°49′2.15″112°47′5.63″2004[58]
SO50-29KL18°26′4.80″115°39′13.20″3766[58]
83PC17°39′31.32″112°32′37.32″1917[93]
SO50-37KL18°54′0″115°48′0″2695[94]
S08-5710°37′1.48″113°42′39.99″3400[90]
V36-319°0′30″116°5′36″2809[58]
V36-619°46′30.00″115°48′30″1597[58]
MD01-239310°30′110°2′60″1230[95]
MD97214212°42′119°28′1557[96]
ODP114420°03′117°25′2036[70]
ODP114309°22′113°71′2772[70]
ODP114818°50′116°34′3297[70]
SO1795613°51′112°35′3387[70]
TP869°23′24″115°40′12″1722[97]
HYD23512°24′36″118°21′2695[97]
ZJ8315°46′48″112°32′24″1511[97]
111PC18°10′12″112°1′12″2253[97]
ZSQD618°30′114°10′48″3020[97]
ZSQD28919°52′12″119°52′12″3605[97]
SCS-15A10°25′114°14′1812[60]
SO49-14KL18°18′114°24′3624[60]
1795710°54′115°18′2195[60]
GIK17937-219.50°117.67°3428[98]
ODP114519.58°117.63°3175[99,100]
191PC19.05°116.22°2510[101]
MD05-290419.46°116.25°2066[102,103]
MD05-290319.46°116.25°2047[104]
MD05-290520°08.2′117°21.6′1647[105]
ZHS17620°00.0′115°33.3′1383[106]
1794020°07.0′117°23.0′1727[107]
MD97214212°41.1′119°27.9′1557[60]
MD05290114°22.5′110°44.60′1454[108]
ODP114420°03′117°25′2036[70]
MD01-239310°30′110°2′60″1230[95]
SCS118°30′22.56″116°16′5.28″3770This study
Figure 3. The geographic locations of sediment cores were analyzed in this study. Red dots indicate that these cores used for stratigraphic comparison. Yellow dots indicate that these cores are notable for exhibiting an early Holocene low-carbonate event, in contrast to core SCS1. For detailed locations and references, see Table 1. The HydroRIVERS dataset [109] can be accessed at www.hydrosheds.org (accessed on 2 January 2024).
Figure 3. The geographic locations of sediment cores were analyzed in this study. Red dots indicate that these cores used for stratigraphic comparison. Yellow dots indicate that these cores are notable for exhibiting an early Holocene low-carbonate event, in contrast to core SCS1. For detailed locations and references, see Table 1. The HydroRIVERS dataset [109] can be accessed at www.hydrosheds.org (accessed on 2 January 2024).
Jmse 12 00438 g003
Figure 4. Downcore variations in carbonate percentage across 23 extended cores within the SCS. For specific locations and references, please refer to Table 1. The numerals following “SB” denote carbonate stratigraphic events.
Figure 4. Downcore variations in carbonate percentage across 23 extended cores within the SCS. For specific locations and references, please refer to Table 1. The numerals following “SB” denote carbonate stratigraphic events.
Jmse 12 00438 g004

4.2. Astronomical Calibration of the Time Scale

4.2.1. Initial Age Model and Spectral Analysis

An initial age–depth model for core SCS1 sediments was established using carbonate stratigraphy (Figure 5). Ages were interpolated and extrapolated across samples by assuming uniform sedimentation rates from the core’s top (considered to be of present age) to all established age control points. Linear sedimentation rates for various stages were calculated based on the age–depth relationship (Figure 5), suggesting the base of core SCS1 to be approximately 100 kyr old, with an overall average sedimentation rate of around 8 cm/kyr. This rate aligns with those from other studies in the northern SCS near our core location [21,100,110]. According to the variation in sedimentation rate (Figure 5), we found that the average sedimentation rate of Stage I and II in the carbonate content record is 6.1 cm/kyr, which belongs to the MIS 3–5 period; the average sedimentation rate of stage III is 17.6 cm/kyr, which belongs to the MIS 2 period; and the average sedimentation rate of stage IV is 8.1 cm/kyr, which belongs to the MIS 1 period. The sedimentation rate during the glacial period is higher than that during the interglacial period, with the highest sedimentation rate during the MIS 2 period in the last glacial period.
We used the statistical software program Acycle v2.4.1 [86] to conduct spectral analyses on the carbonate datasets on the initial age–depth model. The power spectra demonstrated that the carbonate record is dominated by 20 kyr precession paced cycles (Figure 6C). Based on the initial age model, the spectral analysis of the carbonate data series shows that the sedimentary variations in core SCS1 correlate with Earth’s orbital changes, marking an orbital-scale climate cycle. This is a crucial basis for employing the precession series as target curves for refining the initial age scales. The detection of the precession cycle implies low-latitude process influences on core SCS1 deposition, with confidence levels for these climate cycles exceeding 99%, indicating reliable spectral analysis.

4.2.2. Astronomical Calibration

An initial age model was developed by aligning the carbonate curve of core SCS1 with the established carbonate stratigraphic time scale of the SCS. However, this initial model, rooted in carbonate stratigraphy, does not suffice for high-resolution studies that scrutinize the climate’s response to orbital forces. A more direct approach involves tuning the carbonate record to Earth’s orbital parameters, a method that not only aligns with the establishment of astronomical time scales but also promises enhanced age precision. The carbonate sequence of core SCS1 was adjusted to match the corresponding orbital parameters, laying the groundwork for an orbital-scale chronological framework. Analysis of the carbonate sequence, based on the initial age model, reveals that precession cycles prominently influence the sedimentary record (Figure 6C). Consequently, we selected Earth’s orbital precession as our astronomical target curve, employing the La2004 precession solution [87] as our reference for tuning. The La2004 orbital solution [87] is calculated using modern input values for the dynamical ellipticity of the Earth and the tidal dissipation in the evolution of the Earth–Moon system. This orbital solution has been improved with respect to La1993 and proved to have an accurate solution for the geological records [2,47,51].
The carbonate record consistently captures the 20 kyr precession cycle throughout the study period, as depicted in Figure 6B, and is thus chosen as the tuning signal. Leveraging the initial age model, we embark on a detailed calibration of the carbonate record to the 20 kyr precession cycle. Prior research suggests that carbonate content is likely a direct regional response to insolation forcing [56], leading us to base our tuning on the assumption of a steady and linear in-phase relationship between the tuning signal and the precession target. In their study, Zeeden et al. [111] tuned the magnetic susceptibility (MS) maxima to precession maxima in the astronomical tuning of core ODP 926. The low carbonate contents usually induce high MS values in the SCS [2,51]. Thus, the high values of MS indicate a relatively low carbonate content, since it is inversely correlated to the carbonate record [54,111,112]. Accordingly, we associate the lowest carbonate content with the highest precession values. Furthermore, Li’s [108] analysis further refines this relationship by demonstrating that peaks in carbonate content align with troughs in precession values, directly linking carbonate maxima to precession minima in the SCS cores. Thus, the phase relationship between carbonate cycles and precession parameters is characterized by a correspondence where the lowest (highest) carbonate levels coincide with the precession index’s peak (trough) values. Precession dominates productivity and Asian winter monsoon changes in the SCS [66,67,68,113,114,115]. Meanwhile, the productivity change in the northern SCS is affected by the Asian winter monsoon [114,116,117]. The minimum value of the precession causes the enhancement of the Asian winter monsoon, which leads to the enhancement of carbonate productivity, and the opposite is true when the precession is at the maximum value. A detailed discussion is presented in Section 5.2.
We tuned our 20 kyr cycle of the carbonate record to the 20 kyr cycle component curve extracted from the precession to develop an astronomical time scale for core SCS1 [87]. Our analysis identified approximately five and a half 20 kyr cycles within the filtered carbonate content (Figure 6A). The precession model for the 0 to 120 kyr interval revealed six 20 kyr cycles (Figure 6D), aligning closely with observations from visual assessments and Gaussian filtering. We manually aligned the carbonate maxima to 20 kyr precession minima by visually selecting tie points (Figure 6) in the QAnalySeries v1.5.0 software [88]. By iteratively refining the initial age model with precession data, we achieved maximum coherence between the filtered carbonate and orbital signals, resulting in a high-resolution time scale. The congruence between the carbonate record’s 20 kyr filter and the precession curve’s 20 kyr component (Figure 6A,D) underscores the reliability of our age model. The tuned carbonate time series date was used to perform a power spectral analysis in Acycle v2.4.1 software. The method was the multi-taper method (MTM) and robust AR (1) red noise model. The power spectrum of the tuned carbonate time series indicates spectral peaks at periodicities of 20 kyr (Figure 6I). We also used Acycle v2.4.1 software to carry out evolutionary spectral analysis of tuned carbonate time series date by fast Fourier transform. In the evolutionary FFT spectrogram (Figure 6H), one can observe persistent power in the 20 kyr band. The core SCS1 records, as per the carbonate-tuned age model, span from 0 to approximately 110 kyr (Figure 6G), with evolutionary spectral analyses of the tuned carbonate time series revealing a pronounced expression of the precession cycle within the dataset (Figure 6H).
The validity and accuracy of the established astronomical time scale can be gauged through time-domain spectral analysis. In the time-domain spectral analysis, if the signals of the Earth’s orbital parameter cycles used for astronomical tuning are strong and the peaks of the other parameter cycles are within the confidence interval, the established astronomical age scale is more reliable [118]. The spectral analysis results after tuning are consistent with those of the spectral analysis of the initial age model. Time-domain spectral analysis of the tuned carbonate series (Figure 6I) confirms the dominance of a 20 kyr precession cycle, with confidence levels for all cycles exceeding 99%. These results corroborate the presence of the expected astronomical cycles, demonstrating the accuracy of both the astronomical cycle signal interpretation and the established astronomical time scale.

5. Discussion

5.1. Low-Carbonate Event

The carbonate record from core SCS1 spans marine isotope stage (MIS) 1 to 5. The average carbonate content is notably low at 13.6%, except for a significant increase during MIS 3 (Figure 6G). Peaks in the carbonate curve are evident in MIS 3 and late MIS 2, whereas the lowest values are observed in MIS 4, 5, and the late phases of MIS 1 and 3. Additionally, the carbonate content experiences a peak in late MIS 2 before declining towards MIS 1 and 3 (Figure 6G). The carbonate content curve largely reflects the “Pacific-type” carbonate cycle [60]. That is, both MIS 5 and 4 record a significantly lower carbonate content, which is a common feature of the equatorial Pacific Ocean; MIS 3 records a high carbonate content, which is consistent with the higher carbonate content in MIS 3 than in MIS 2 of the “Pacific-type” cycle [58,119]. Similar carbonate cycles are observed not only in core SO-50-29 from the northern SCS but also in core TT013-PC72 from the equatorial Pacific [58,120]. Anomalies in carbonate content are particularly notable during the Last Glacial Period within core SCS1. MIS 3, an interglacial step within this period, is characterized by a high carbonate content, while the glacial step (MIS 4 and 2) exhibits relatively lower levels. The Last Glacial Maximum, in particular, recorded the lowest carbonate content within this period. The Holocene’s average carbonate content, at 11.3%, falls below that of the Last Glacial Period. It can be seen that core SCS1’s carbonate record not only follows the overall “Pacific-type” cyclic pattern but also exhibits “Atlantic-type” cyclic characteristics during the Last Glacial Period.
The carbonate content increased from the Last Glacial Period to the Holocene [62,121]. However, an abrupt decrease in carbonate content during the early Holocene is noted, with the content dropping from 20.1% to 7.9%, a 61% reduction (Figure 7A). Following this decrease, the sediment’s carbonate content gradually recovers (Figure 7). This phenomenon is termed the early Holocene low-carbonate event (~11–8 kyr) [62,89,98,106,121,122]. It is widespread in the sedimentary cores of the broad continental slope of the northern SCS. It occurs relatively simultaneously in many sedimentary cores [62], for example, the Xisha Trough [89], the Pearl River Canyon [122], and near the Dongsha Islands [105]. To mitigate the influence of regional characteristics, 13 cores of relatively high-resolution records were selected for comparative analysis (Table 1; Figure 3 and Figure 7), revealing that the early Holocene low-carbonate event is common across both deep and shallow water cores, with consistent results regarding its timing and duration within dating uncertainties (Figure 7). These comparisons further validate the reliability of the astronomical time scale based on carbonate stratigraphy at core SCS1.
The cause of the early Holocene low-carbonate event in core SCS1 prompts investigation. Considering that changes in seawater chemistry impact the carbonate system throughout the SCS, yet no similar event is observed in the western and southern parts of the sea, the hypothesis of enhanced carbonate dissolution is discounted [62]. This event is likely related to dilution, coinciding with increased rainfall from the East Asian summer monsoon, as indicated by sedimentary records from Dongge Cave stalagmites and Lake Huguang Maar (Figure 7I,R), showing peak rainfall during ~11–8.0 kyr [123,124,125]. This increase in rainfall could have intensified erosion in northern Hainan Island and southern China [89,126,127], thereby enhancing river unloading capacity and the dilution of deep-sea carbonate with terrestrial materials. Additionally, sudden cooling [128,129] and volcanic events [58,130] could have further increased terrestrial material input, diluting the carbonate content. Huang et al. [62] attributed the low-carbonate event to sediment dilution from Taiwan and Luzon, a finding further supported by Luo et al. [121]. Besides these factors, a decrease in surface ocean productivity is another potential contributor [98,122]. The δ13Corg values in the core, indicating organic-matter sources in sediments [131], show a significant decrease during this period (Figure 7P), alongside a notable reduction in total organic carbon (TOC) content (Figure 7Q). This suggests a decline in surface seawater productivity, impacting the carbonate content. Consequently, the early Holocene low-carbonate event in core SCS1 is attributed to terrestrial material dilution, coupled with reduced surface seawater productivity.
Figure 7. The variability in carbonate content and other relevant proxies around the SCS to illustrate the early Holocene low-carbonate event: (A,J) Display the carbonate content curves for core SCS1. (P,Q) Present the total organic carbon (TOC) and δ13Corg curves for core SCS1, respectively. (BH,KO) Feature carbonate content curves from various cores, including GIK17937-2, ODP1145, 191PC, MD05-2904, MD05-2903, ODP1144, 17940, MD05-2905, MD97-2142, MD052901, ZHS176, and MD01-2393. Locations and references for these cores are detailed in Table 1. Comparison of low-carbonate event around the SCS. (I) Shows the titanium (Ti) content from the sediment sequence of Lake Huguang Maar, indicating changes in environmental conditions [123]. (R) Represents oxygen isotope values from stalagmites in Dongge Cave, offering insights into past climatic fluctuations [125]. A gray shaded bar across the figure highlights the period of the early Holocene low-carbonate event, visually correlating the data points and illustrating the widespread nature of this event across different locations in the SCS. This comprehensive comparison helps to further understand the timing, extent, and potential causes of the low-carbonate event during the early Holocene.
Figure 7. The variability in carbonate content and other relevant proxies around the SCS to illustrate the early Holocene low-carbonate event: (A,J) Display the carbonate content curves for core SCS1. (P,Q) Present the total organic carbon (TOC) and δ13Corg curves for core SCS1, respectively. (BH,KO) Feature carbonate content curves from various cores, including GIK17937-2, ODP1145, 191PC, MD05-2904, MD05-2903, ODP1144, 17940, MD05-2905, MD97-2142, MD052901, ZHS176, and MD01-2393. Locations and references for these cores are detailed in Table 1. Comparison of low-carbonate event around the SCS. (I) Shows the titanium (Ti) content from the sediment sequence of Lake Huguang Maar, indicating changes in environmental conditions [123]. (R) Represents oxygen isotope values from stalagmites in Dongge Cave, offering insights into past climatic fluctuations [125]. A gray shaded bar across the figure highlights the period of the early Holocene low-carbonate event, visually correlating the data points and illustrating the widespread nature of this event across different locations in the SCS. This comprehensive comparison helps to further understand the timing, extent, and potential causes of the low-carbonate event during the early Holocene.
Jmse 12 00438 g007

5.2. Effects of Precessional Forcing on Carbonate Content Variations in the Low-Latitude Sea

Since the Quaternary, the formation and expansion of the Northern Hemisphere ice sheet have played a dominant role in the global climate. However, as an important part of the global climate system, low latitudes also play an essential role. The apparent precession cycle in the carbonate record of core SCS1 is a typical low-latitude feature. It should be the response of low-latitude processes (such as the East Asian monsoon) to orbital parameters. Thus, variations in carbonate content in core SCS1 respond to precession forcing, i.e., driven by low-latitude processes.
The precession curve of the tuned carbonate time series is consistent with that of the northern hemisphere summer solar insolation (Figure 6E,F). It shows that the change in carbonate content is influenced by solar insolation. The greater the solar insolation, the higher the temperature, and thus, the more carbonate is formed. Moreover, spectral analysis underscores the dominant role of precession forcing in modulating productivity changes within the SCS region [113]. The evident strong precession cycle in core SCS1’s carbonate record suggests that productivity changes significantly affect carbonate preservation at this site [114]. The variability in surface ocean productivity in the SCS, particularly in its northern parts, is closely linked to the intensity of the Asian winter monsoon systems [114,116,117]. Sedimentary records from the SCS further reveal that these monsoon systems are largely influenced by the precession cycle [66,67,68,115]. Analysis of insolation changes induced by orbital parameters over the last 300,000 years [65,132] shows that minimum precession coincides with maximal northern hemisphere summer insolation and minimal winter insolation, and vice versa for maximum precession [69]. When the northern hemisphere winter insolation caused by precession is weakened (enhanced), the temperature decrease (increase) in the southern East Asian continent is significantly greater than the temperature change in the neighboring western Pacific region, thus strengthening (weakening) the latitudinal land–sea thermal contrast, and then strengthening (weakening) the winter monsoon in the southern East Asia region [67,133]. Thus, when the precession is at a minimum, the Northern Hemisphere winter insolation is low, the East Asian winter monsoon is stronger [66,67,115].
Intensification of the East Asian winter monsoon boosts surface ocean productivity through several mechanisms [6,108,113,134,135,136]. On the one hand, the enhanced winter monsoon increases the input of nutrients in dust deposition, increasing the surface seawater productivity [117,137]. On the other hand, the enhanced winter monsoon drives the surface seawater to flow southward, the thermocline in the northern SCS becomes shallow, the subsurface eutrophic water enters the euphotic layer, and the surface productivity increases. At the same time, the intense winter monsoon can accelerate the vertical circulation of seawater and promote stronger vertical mixing. This brings the trophic cold water below the mixed layer to the surface of the seawater to form a wind-driven upwelling. The upwelling is strong, and the deep water rich in nutrients rises to the surface water, which increases productivity [6,134,136]. In addition, the winter monsoon promotes the North Pacific current to flow through the Taiwan Strait and the Bashi Strait into the SCS, forming a coastal current and bringing more nutrients, resulting in increased surface productivity [138]. Furthermore, the strong East Asian winter monsoon is conducive to the seasonal Kuroshio Current (KC) or Taiwan Warm Current flowing into the northern SCS [139,140]; when the KC flows into the SCS, its nutrient-rich deep water is carried to the surface by upwelling driven by different processes, including wind-driven upwelling, vertical mixing, or cyclonic vortices [141,142], further enhancing productivity. The increase in biological productivity increases the absolute deposition of carbonate, the deepening of the lysocline of carbonate, and the decrease in solubility. Thus, higher productivity is conducive to the preservation of carbonate and increases the carbonate content. The conspicuous precession cycle in our records indicates that the positive impact of increased biological productivity on carbonate deposition outweighs the negative effects of other factors. This analysis suggests that a strong winter monsoon enhances surface seawater productivity during minimum precession phases, favoring carbonate deposition and preservation; the opposite occurs during maximum precession phases. It is noteworthy that monsoon dynamics can also be viewed as manifestations of the seasonal migration of the Intertropical Convergence Zone (ITCZ), with the precession cycle influencing the global monsoon through north–south ITCZ displacement [70,113,143]. Consequently, in the precession band, carbonate deposition and preservation in core SCS1 are influenced by the interplay of the ITCZ and monsoon systems.

6. Conclusions

In this study, we have detailed a high-resolution carbonate record from core SCS1, collected from the depths of the SCS. By integrating carbonate stratigraphy with time series analyses of the carbonate record, we have successfully established an age-calibrated astronomical time scale for core SCS1. The initial analysis of an untuned carbonate stratigraphic age model revealed that precession is the most prominently expressed astronomical parameter. Consequently, an astronomical time scale covering the past ~110 kyr was developed by aligning the time-domain carbonate data with the La2004 precession curve. This astronomical time scale reveals that the carbonate record of core SCS1 exhibits a “Pacific-type” carbonate cycle. Moreover, the timing and duration of the low-carbonate event identified in core SCS1 align with findings from previous research. This event is primarily attributed to dilution effects coupled with a reduction in productivity. The pronounced precession cycle observed in the carbonate record of core SCS1 underscores that changes in carbonate content are driven by precessional forcing, indicative of low-latitude processes. Within the precession band, the interplay between the Intertropical Convergence Zone (ITCZ) and the monsoon system significantly influences carbonate deposition and preservation by modulating productivity variations in the SCS. The insights gained from this study significantly enhance our understanding of paleoceanographic and paleoclimatic evolution in the northern SCS, highlighting the intricate relationship between astronomical parameters and marine sedimentation processes. However, this study did not reconstruct the paleoclimate evolution of the northern SCS, so in future studies, we can combine a variety of climate proxy indicators to carry out research on the paleoclimate evolution of the northern SCS.

Author Contributions

C.Z. (Chunhui Zhang) participated in the whole work and drafting of the manuscript; W.Z. and C.Z. (Chengjun Zhang) contributed to conception, design, and final approval of the version to be published; L.Z., S.Y., Y.M. and W.D. contributed to data analysis. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Second Tibetan Plateau Scientific Expedition and Research Program (STEP), Grant No. 2019QZKK0704.

Data Availability Statement

The original contributions presented in this study are included in the article; further inquiries can be directed to the corresponding author.

Acknowledgments

We thank researcher Weiguo Wang (Third Institute of Oceanography, Ministry of Natural Resources, Xiamen 361005, China) for providing the columnar samples of marine sediments studied in this paper.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Tian, J.; Wang, P.X.; Cheng, X.R.; Li, Q.Y. Astronomically tuned Plio–Pleistocene benthic δ18O record from South China Sea and Atlantic–Pacific comparison. Earth Planet. Sci. Lett. 2002, 203, 1015–1029. [Google Scholar] [CrossRef]
  2. Tian, J.; Zhao, Q.H.; Wang, P.X.; Li, Q.Y.; Cheng, X.R. Astronomically modulated Neogene sediment records from the South China Sea. Paleoceanography 2008, 23, PA3210. [Google Scholar] [CrossRef]
  3. Wang, L.; Li, G.; Liu, J.; Mei, X.; Zhang, Y. Astronomical dating of Quaternary strata in the South Yellow Sea and its indication for paleoclimatic evolution. Mar. Geol. 2021, 439, 106557. [Google Scholar] [CrossRef]
  4. Nie, Y.; Wu, H.; Satolli, S.; Ferré, E.C.; Shi, M.; Fang, Q.; Xu, Y.; Zhang, S.; Li, H.; Yang, T. Late Miocene to present paleoclimatic and paleoenvironmental evolution of the South China Sea recorded in the Magneto-Cyclostratigraphy of IODP Site U1505. Paleoceanogr. Paleoclimatol. 2023, 38, e2022PA004547. [Google Scholar] [CrossRef]
  5. Wang, P.X.; Li, Q.Y.; Tian, J. Pleistocene paleoceanography of the South China Sea: Progress over the past 20 years. Mar. Geol. 2014, 352, 381–396. [Google Scholar] [CrossRef]
  6. Tian, J.; Wang, P.; Chen, R.; Cheng, X. Quaternary upper ocean thermal gradient variations in the South China Sea: Implications for east Asian monsoon climate. Paleoceanography 2005, 20, PA4007. [Google Scholar] [CrossRef]
  7. Liu, Z.; Alain, T.; Clemens, S.C.; Wang, P. Quaternary clay mineralogy in the northern South China Sea (ODP Site 1146). Sci. China Ser. D Earth Sci. 2003, 46, 1223–1235. [Google Scholar] [CrossRef]
  8. Zhang, C.; Yang, S.Y.; Huang, X.T.; Dou, Y.G.; Li, F.L.; Xu, X.N.; Hao, Q.; Gao, J.H. Sea level change and Kuroshio intrusion dominated Taiwan sediment source-to-sink processes in the northeastern South China Sea over the past 244 kyrs. Quat. Sci. Rev. 2022, 287, 107558. [Google Scholar] [CrossRef]
  9. Liu, Z.; Chen, X.; Yu, J.; Xu, D.; Sun, C. Kuroshio intrusion into the South China Sea with an anticyclonic eddy: Evidence from underwater glider observation. J. Oceanol. Limnol. 2019, 37, 1469–1480. [Google Scholar] [CrossRef]
  10. Hu, D.; Wu, L.; Cai, W.; Gupta, A.S.; Ganachaud, A.; Qiu, B.; Gordon, A.L.; Lin, X.; Chen, Z.; Hu, S.; et al. Pacific western boundary currents and their roles in climate. Nature 2015, 522, 299–308. [Google Scholar] [CrossRef]
  11. Zhao, M.; Huang, C.-Y.; Wang, C.-C.; Wei, G. A millennial-scale U 37 K sea-surface temperature record from the South China Sea (8° N) over the last 150 kyr: Monsoon and sea-level influence. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2006, 236, 39–55. [Google Scholar] [CrossRef]
  12. Chen, L.; Chen, L.; Yin, Z.X.; Guan, Y.L.; Zhang, Y.Z.; Li, G.S.; Jiang, Z.X. Magnetic characteristics of sediments from the central basin of the South China Sea since the late Pleistocene: Implications for sediment provenance and evolution of the East Asian monsoon. Chin. J. Geophys. 2023, 66, 3557–3575, (In Chinese with English Abstract). [Google Scholar]
  13. Liu, Z.F.; Zhao, Y.L.; Colin, C.; Stattegger, K.; Wiesner, M.G.; Huh, C.A.; Zhang, Y.W.; Li, X.J.; Sompongchaiyakul, P.; You, C.F.; et al. Source-to-sink transport processes of fluvial sediments in the South China Sea. Earth-Sci. Rev. 2016, 153, 238–273. [Google Scholar] [CrossRef]
  14. Li, W.; Li, X.; Mei, X.; Zhang, F.; Xu, J.; Liu, C.; Wei, C.; Liu, Q. A review of current and emerging approaches for Quaternary marine sediment dating. Sci. Total Environ. 2021, 780, 146522. [Google Scholar] [CrossRef] [PubMed]
  15. Lin, P.; Song, Y.; Zhan, W.; Tian, R.; Wang, Z.; Xu, X.; Luo, L.; Abbas, M.; Lai, Z. Late Pleistocene to Holocene sedimentary history in the Pearl River Delta revealed by OSL and radiocarbon dating. Catena 2023, 224, 106972. [Google Scholar] [CrossRef]
  16. Zhong, J.; Ling, K.; Yang, M.; Shen, Q.; Abbas, M.; Lai, Z. Radiocarbon and OSL dating on cores from the Chaoshan delta in the coastal South China Sea. Front. Mar. Sci. 2022, 9, 1030841. [Google Scholar] [CrossRef]
  17. Zhao, H.; Lu, Y.; Wang, C.; Mao, H.; Ji, H. A review of OSL dating for water-laid deposits: Progress and prospect. Nucl. Technol. 2011, 34, 82–86. [Google Scholar]
  18. Shen, X.; Hu, B.; Yan, H.; Dodson, J.; Zhao, J.; Li, J.; Ding, X.; Li, Q.; Wang, X.; Xu, F. Reconstruction of Kuroshio intrusion into the South China Sea over the last 40 kyr. Quat. Sci. Rev. 2022, 290, 107622. [Google Scholar] [CrossRef]
  19. Li, M.K.; Ouyang, T.P.; Tian, C.J.; Zhu, Z.Y.; Peng, S.S.; Tang, Z.H.; Qiu, Y.; Zhong, H.X.; Peng, X.C. Sedimentary responses to the East Asian monsoon and sea level variations recorded in the northern South China Sea over the past 36 kyr. J. Asian Earth Sci. 2019, 171, 213–224. [Google Scholar] [CrossRef]
  20. Li, M.K.; Ouyang, T.P.; Zhu, Z.Y.; Tian, C.J.; Peng, S.S.; Zhong, H.X.; Peng, X.C.; Qiu, Y. Reconstruction of chemical weathering intensity and Asian summer monsoon evolution in the Red River basin over the past 36 kyr. Paleoceanogr. Paleoclimatol. 2022, 37, e2021PA004397. [Google Scholar] [CrossRef]
  21. Xu, F.J.; Hu, B.Q.; Zhao, J.T.; Liu, X.T.; Xu, K.H.; Xiong, Z.F.; Wang, F.F.; Ding, X.; Li, Q.; Guo, J.W. Provenance and weathering of sediments in the deep basin of the northern South China Sea during the last 38 kyr. Mar. Geol. 2021, 440, 106602. [Google Scholar] [CrossRef]
  22. Reimer, P.J.; Bard, E.; Bayliss, A.; Beck, J.W.; Blackwell, P.G.; Ramsey, C.B.; Buck, C.E.; Cheng, H.; Edwards, R.L.; Friedrich, M.; et al. IntCal13 and Marine13 Radiocarbon Age Calibration Curves 0–50,000 Years cal BP. Radiocarbon 2013, 55, 1869–1887. [Google Scholar] [CrossRef]
  23. Broecker, W.; Barker, S.; Clark, E.; Hajdas, I.; Bonani, G. Anomalous radiocarbon ages for foraminifera shells. Paleoceanography 2006, 21, PA2008. [Google Scholar] [CrossRef]
  24. Bard, E.; Arnold, M.; Duprat, J.; Moyes, J.; Duplessy, J.C. Reconstruction of the last deglaciation: Deconvolved records of δ18O profiles, micropaleontological variations and accelerator mass spectrometric14C dating. Clim. Dyn. 1987, 1, 101–112. [Google Scholar] [CrossRef]
  25. Barker, S.; Broecker, W.; Clark, E.; Hajdas, I. Radiocarbon age offsets of foraminifera resulting from differential dissolution and fragmentation within the sedimentary bioturbated zone. Paleoceanography 2007, 22, PA2205. [Google Scholar] [CrossRef]
  26. Mekik, F. Radiocarbon dating of planktonic foraminifer shells: A cautionary tale. Paleoceanography 2014, 29, 13–29. [Google Scholar] [CrossRef]
  27. Wycech, J.; Kelly, D.C.; Marcott, S. Effects of seafloor diagenesis on planktic foraminiferal radiocarbon ages. Geology 2016, 44, 551–554. [Google Scholar] [CrossRef]
  28. Missiaen, L.; Waelbroeck, C.; Pichat, S.; Jaccard, S.L.; Eynaud, F.; Greenop, R.; Burke, A. Improving North Atlantic marine core chronologies using 230Th normalization. Paleoceanogr. Paleoclimatol. 2019, 34, 1057–1073. [Google Scholar] [CrossRef]
  29. Lisiecki, L.E.; Raymo, M.E. A Pliocene-Pleistocene stack of 57 globally distributed benthic δ18O records. Paleoceanography 2005, 20, PA1003. [Google Scholar]
  30. Geibert, W.; Stimac, I.; Rutgers van der Loeff, M.M.; Kuhn, G. Dating deep-sea sediments with 230Th excess using a constant rate of supply model. Paleoceanogr. Paleoclimatol. 2019, 34, 1895–1912. [Google Scholar] [CrossRef]
  31. Wu, H.; Shi, M.; Zhao, X.; Huang, B.; Zhang, S.; Li, H.; Yang, T.; Lin, C. Magnetostratigraphy of ODP Site 1143 in the South China Sea since the Early Pliocene. Mar. Geol. 2017, 394, 133–142. [Google Scholar] [CrossRef]
  32. Yang, X.Q.; Friedrich, H.; Wu, N.Y.; Yang, J.; Su, Z.H. Geomagnetic paleointensity dating of South China Sea sediments for the last 130 kyr. Earth Planet. Sci. Lett. 2009, 284, 258–266. [Google Scholar]
  33. Pillans, B.; Gibbard, P. Chapter 30—The Quaternary Period. In The Geologic Time Scale; Gradstein, F.M., Ogg, J.G., Schmitz, M.D., Ogg, G.M., Eds.; Elsevier: Boston, MA, USA, 2012; pp. 979–1010. [Google Scholar]
  34. Sakuramoto, Y.; Yamazaki, T.; Kimoto, K.; Miyairi, Y.; Kuroda, J.; Yokoyama, Y.; Matsuzaki, H. A geomagnetic paleointensity record of 0.6 to 3.2 Ma from sediments in the western equatorial Pacific and remanent magnetization lock-in depth. J. Geophys. Res. Solid Earth 2017, 122, 7525–7543. [Google Scholar] [CrossRef]
  35. Simon, Q.; St-Onge, G.; Hillaire-Marcel, C. Late Quaternary chronostratigraphic framework of deep Baffin Bay glaciomarine sediments from high-resolution paleomagnetic data. Geochem. Geophys. Geosyst. 2012, 13, Q0AO03. [Google Scholar] [CrossRef]
  36. Jin, H.; Wan, S.; Clift, P.D.; Liu, C.; Huang, J.; Jiang, S.; Li, M.; Qin, L.; Shi, X.; Li, A. Birth of the Pearl River at 30 Ma: Evidence from sedimentary records in the northern South China Sea. Earth Planet. Sci. Lett. 2022, 600, 117872. [Google Scholar] [CrossRef]
  37. Yu, S.; Huang, C.; Zhang, R.; Wang, Z.; Ogg, J.; Kemp, D.B. Astronomical time scale for the lower Doushantuo Formation of early Ediacaran, South China. Sci. Bull. 2018, 63, 1485–1494. [Google Scholar]
  38. Huang, H.; Gao, Y.; Ma, C.; Jones, M.; Zeeden, C.; Ibarra, D.; Wu, H.; Wang, C. Organic carbon burial is paced by a ~173-ka obliquity cycle in the middle to high latitudes. Sci. Adv. 2021, 7, 9489–9498. [Google Scholar] [CrossRef]
  39. Xu, K.; De Vleeschouwer, D.; Vahlenkamp, M.; Yang, R.; Chen, H. Reconstructing eocene eastern Indian Ocean dynamics using ocean-drilling stratigraphic records. Paleoceanogr. Paleoclimatology 2021, 36, e2020PA004116. [Google Scholar] [CrossRef]
  40. Chen, G.; Gang, W.; Tang, H.; Gao, G.; Wang, N.; Liu, L.; Yang, S.; Wang, Y. Astronomical cycles and variations in sediment accumulation rate of the terrestrial lower Cretaceous Xiagou Formation from the Jiuquan Basin, NW China. Cretac. Res. 2020, 109, 104156. [Google Scholar] [CrossRef]
  41. Sun, Y.; Clemens, S.C.; An, Z.; Yu, Z. Astronomical timescale and palaeoclimatic implication of stacked 3.6 Myr monsoon records from the Chinese Loess Plateau. Quat. Sci. Rev. 2006, 25, 33–48. [Google Scholar] [CrossRef]
  42. Milankovitch, M. Théorie Mathématique des Phénomènes Thermiques Produits Par la Radiation Solaire; Gauthier-Villars et Cie: Paris, France, 1920. [Google Scholar]
  43. Milankovitch, M.K. Kanon der Erdbestrahlung und seine Anwendung auf das Eiszeitenproblem. R. Serbian Acad. Spec. Publ. 1941, 133, 1–633. [Google Scholar]
  44. Hays, J.D.; Imbrie, J.; Shackleton, N.J. Variations in the Earth’s Orbit: Pacemaker of the Ice Ages. Science 1976, 194, 1121–1132. [Google Scholar] [CrossRef]
  45. Yu, Z.W.; Ding, Z.L. An automatic orbital tuning method for paleoclimate records. Geophys. Res. Lett. 1998, 25, 4525–4528. [Google Scholar] [CrossRef]
  46. Zeeden, C.; Meyers, S.R.; Hilgen, F.J.; Lourens, L.J.; Laskar, J. Time scale evaluation and the quantification of obliquity forcing. Quat. Sci. Rev. 2019, 209, 100–113. [Google Scholar] [CrossRef]
  47. Ao, H.; Dekkers, M.J.; Qin, L.; Xiao, G.Q. An updated astronomical timescale for the Plio-Pleistocene deposits from South China Sea and new insights into Asian monsoon evolution. Quat. Sci. Rev. 2011, 30, 1560–1575. [Google Scholar] [CrossRef]
  48. Meyers, S.R. Cyclostratigraphy and the problem of astrochronologic testing. Earth-Sci. Rev. 2019, 190, 190–223. [Google Scholar] [CrossRef]
  49. Shackleton, N.J.; Berger, A.; Peltier, W.R. An alternative astronomical calibration of the lower Pleistocene timescale based on ODP Site 677. Earth Environ. Sci. Trans. R. Soc. Edinb. 1990, 81, 251–261. [Google Scholar] [CrossRef]
  50. Imbrie, J.; Hays, J.; Martinson, D.; McIntyre, A.; Mix, A.; Morley, J.; Pisias, N.; Prell, W.; Shackleton, N. The Orbital Theory of Pleistocene Climate: Support Frim a Revised Chronology of the Marine δ18O Record; Reidel Publishing Company: Dordrecht, The Netherlands, 1984; Volume 126, pp. 269–305. [Google Scholar]
  51. Zhang, Y.; Yi, L.; Ogg, J.G. Pliocene-Pleistocene magneto-cyclostratigraphy of IODP Site U1499 and implications for climate-driven sedimentation in the northern South China Sea. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2019, 527, 118–132. [Google Scholar] [CrossRef]
  52. Grützner, J.; Giosan, L.; Franz, S.O.; Tiedemann, R.; Cortijo, E.; Chaisson, W.P.; Flood, R.D.; Hagen, S.; Keigwin, L.D.; Poli, S.; et al. Astronomical age models for Pleistocene drift sediments from the western North Atlantic (ODP Sites 1055–1063). Mar. Geol. 2002, 189, 5–23. [Google Scholar] [CrossRef]
  53. Feng, X.G.; Jiang, F.Q.; Zhang, Z.H.; Xiong, Z.F.; Zhong, Y.; Dong, J.; Chen, T.Y.; Li, A.C.; Zou, X.Q.; Shi, X.F. Long eccentricity forcing Asian dust input into the northwestern Pacific during the early Pleistocene. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2022, 596, 110963. [Google Scholar] [CrossRef]
  54. Beddow, H.M.; Liebrand, D.; Wilson, D.S.; Hilgen, F.J.; Sluijs, A.; Wade, B.S.; Lourens, L.J. Astronomical tunings of the Oligocene–Miocene transition from Pacific Ocean Site U1334 and implications for the carbon cycle. Clim. Past 2018, 14, 255–270. [Google Scholar] [CrossRef]
  55. Wehausen, R.; Brumsack, H.J. Astronomical forcing of the East Asian monsoon mirrored by the composition of Pliocene South China Sea sediments. Earth Planet. Sci. Lett. 2002, 201, 621–636. [Google Scholar] [CrossRef]
  56. Liebrand, D.; Beddow, H.M.; Lourens, L.J.; Pälike, H.; Raffi, I.; Bohaty, S.M.; Hilgen, F.J.; Saes, M.J.M.; Wilson, P.A.; van Dijk, A.E.; et al. Cyclostratigraphy and eccentricity tuning of the early Oligocene through early Miocene (30.1–17.1 Ma): Cibicides mundulus stable oxygen and carbon isotope records from Walvis Ridge Site 1264. Earth Planet. Sci. Lett. 2016, 450, 392–405. [Google Scholar] [CrossRef]
  57. Hilgen, F.; Zeeden, C.; Laskar, J. Paleoclimate records reveal elusive ~200-kyr eccentricity cycle for the first time. Glob. Planet. Chang. 2020, 194, 103296. [Google Scholar] [CrossRef]
  58. Qian, J.X. Paleoceanography of the South China Sea since the Late Quaternary; Science Press: Beijing, China, 1999; p. 167. [Google Scholar]
  59. Farrell, J.W.; Prell, W.L. Climatic change and CaCO3 preservation: An 800,000 year bathymetric reconstruction from the central equatorial Pacific Ocean. Paleoceanography 1989, 4, 447–466. [Google Scholar] [CrossRef]
  60. Wang, P.X.; Wang, L.J.; Bian, Y.H.; Jian, Z.M. Late Quaternary paleoceanography of the South China Sea: Surface circulation and carbonate cycles. Mar. Geol. 1995, 127, 145–165. [Google Scholar] [CrossRef]
  61. Huang, C.Y.; Wang, C.C.; Zhao, M.X. High-resolution carbonate stratigraphy of IMAGES core MD972151 from South China Sea. Terr. Atmos. Ocean. Sci. 1999, 10, 225–238. [Google Scholar] [CrossRef]
  62. Huang, E.Q.; Tian, J.; Qiao, P.J.; Wan, S.; Xie, X.; Yang, W.G. Early interglacial carbonate-dilution events in the South China Sea: Implications for strengthened typhoon activities over subtropical East Asia. Quat. Sci. Rev. 2015, 125, 61–77. [Google Scholar] [CrossRef]
  63. Thunell, R.C.; Qingmin, M.; Calvert, S.E.; Pedersen, T.F. Glacial-Holocene biogenic sedimentation patterns in the South China Sea: Productivity variations and surface water pCO2. Paleoceanography 1992, 7, 143–162. [Google Scholar] [CrossRef]
  64. Prell, W.L.; Kutzbach, J.E. Monsoon variability over the past 150,000 years. J. Geophys. Res. Atmos. 1987, 92, 8411–8425. [Google Scholar] [CrossRef]
  65. Berger, A.; Loutre, M.F. Insolation values for the climate of the last 10 million years. Quat. Sci. Rev. 1991, 10, 297–317. [Google Scholar] [CrossRef]
  66. Yu, P.S.; Huang, C.C.; Chin, Y.; Mii, H.S.; Chen, M.T. Late Quaternary East Asian Monsoon variability in the South China Sea: Evidence from planktonic foraminifera faunal and hydrographic gradient records. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2006, 236, 74–90. [Google Scholar] [CrossRef]
  67. Xie, X.; Liu, X.; Chen, G.; Korty, R.L. A transient modeling study of the latitude dependence of East Asian winter monsoon variations on orbital timescales. Geophys. Res. Lett. 2019, 46, 7565–7573. [Google Scholar] [CrossRef]
  68. Huang, E.; Wang, S.; Wei, S.; Yuan, Z.; Tian, J. Precession control of interglacial winter monsoon intensity over tropical East Asia. Glob. Planet. Chang. 2023, 229, 104247. [Google Scholar] [CrossRef]
  69. Liu, X.; Shi, Z. Effect of precession on the Asian summer monsoon evolution: A systematic review. Chin. Sci. Bull. 2009, 54, 3720–3730. [Google Scholar] [CrossRef]
  70. Wang, P.; Li, Q. The South China Sea: Paleoceanography and Sedimentology; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2009; Volume 13. [Google Scholar]
  71. Chu, P.C.; Wang, G.H. Seasonal Variability of Thermohaline Front in the Central South China Sea. J. Oceanogr. 2003, 59, 65–78. [Google Scholar] [CrossRef]
  72. Fang, G.H.; Fang, W.D.; Fang, Y.; Wang, K. A survey of studies on the South China Sea upper ocean circulation. Acta Oceanogr. Taiwanica 1998, 37, 1–16. [Google Scholar]
  73. Yuan, D.L.; Han, W.Q.; Hu, D.X. Surface Kuroshio path in the Luzon Strait area derived from satellite remote sensing data. J. Geophys. Res. Ocean. 2006, 111, C11007. [Google Scholar] [CrossRef]
  74. Qu, T.D.; Girton, J.B.; Whitehead, J.A. Deepwater overflow through Luzon Strait. J. Geophys. Res. Ocean. 2006, 111, C01002. [Google Scholar] [CrossRef]
  75. Webster, P.J. The role of hydrological processes in ocean-atmosphere interactions. Rev. Geophys. 1994, 32, 427–476. [Google Scholar] [CrossRef]
  76. Xia, H.; Li, S.; Shi, M. Three-D numerical simulation of wind-driven current and density current in the Beibu Gulf. Acta Oceanol. Sin. 2001, 20, 455–472. [Google Scholar]
  77. Wu, D.; Wang, Y.; Lin, X.; Yang, J. On the mechanism of the cyclonic circulation in the Gulf of Tonkin in the summer. J. Geophys. Res. Ocean. 2008, 113, C09029. [Google Scholar] [CrossRef]
  78. Caruso, M.J.; Gawarkiewicz, G.G.; Beardsley, R.C. Interannual variability of the Kuroshio intrusion in the South China Sea. J. Oceanogr. 2006, 62, 559–575. [Google Scholar] [CrossRef]
  79. Wang, G.H.; Xie, S.P.; Qu, T.D.; Huang, R.X. Deep South China Sea circulation. Geophys. Res. Lett. 2011, 38, L05601. [Google Scholar] [CrossRef]
  80. Zhao, W.; Zhou, C.; Tian, J.; Yang, Q.; Wang, B.; Xie, L.; Qu, T. Deep water circulation in the Luzon Strait. J. Geophys. Res. Ocean. 2014, 119, 790–804. [Google Scholar] [CrossRef]
  81. Önorm, L. Chemical Analyses of Soils–Determination of Carbonate; Österreichisches Normungsinstitut: Wiena, Austria, 1989. [Google Scholar]
  82. Cleveland, W.S. Robust locally weighted regression and smoothing scatterplots. J. Am. Stat. Assoc. 1979, 74, 829–836. [Google Scholar] [CrossRef]
  83. Thomson, D.J. Spectrum estimation and harmonic analysis. Proc. IEEE 1982, 70, 1055–1096. [Google Scholar] [CrossRef]
  84. Mann, M.E.; Lees, J.M. Robust estimation of background noise and signal detection in climatic time series. Clim. Chang. 1996, 33, 409–445. [Google Scholar] [CrossRef]
  85. Kodama, K.P.; Hinnov, L.A. Rock Magnetic Cyclostratigraphy; John Wiley & Sons: Hoboken, NJ, USA, 2014. [Google Scholar]
  86. Li, M.S.; Hinnov, L.; Kump, L. Acycle: Time-series analysis software for paleoclimate research and education. Comput. Geosci. 2019, 127, 12–22. [Google Scholar] [CrossRef]
  87. Laskar, J.; Robutel, P.; Joutel, F.; Gastineau, M.; Correia, A.; Levrard, B.A. A long-term numerical solution for the insolation quantities of the Earth. Astron. Astrophys. 2004, 428, 261–285. [Google Scholar] [CrossRef]
  88. Kotov, S.; Paelike, H. QAnalySeries-a cross-platform time series tuning and analysis tool. In Proceedings of the AGU Fall Meeting Abstracts, Washington, DC, USA, 10–14 December 2018; p. PP53D-1230. [Google Scholar]
  89. Li, X.J.; Jiang, M.S. Low carbonate event in northern South China Sea during the early Holocene and its paleoclimatic significance. J. Palaeogeogr. 2003, 5, 355–364, (In Chinese with English Abstract). [Google Scholar]
  90. Wang, C.; Gong, Q.J.; Shi, X.F.; She, M.; Wei, X.U. Geochemical characteristics of core S08-57 in South China Sea and their paleoceanographic significance. J. Trop. Oceanogr. 2007, 26, 37–42, (In Chinese with English Abstract). [Google Scholar]
  91. Lu, H.F.; Liu, J.; Chen, F.; Zhou, Y.; Liao, Z.L.; Zhang, X.; Zheng, K.Q. CaCO3 chemostratigraphy of sediment cores from northeastern South China Sea. Gresearch Eological South China Sea 2008, 1, 14–22, (In Chinese with English Abstract). [Google Scholar]
  92. Stuiver, M.; Reimer, P.J. CALIB rev. 8. Radiocarbon 1993, 35, 215–230. [Google Scholar] [CrossRef]
  93. Zhuang, C.; Chen, F.; Jinpeng, Z.; Yang, Z.; Sihai, C. Planktonic foraminifera in the Xisha Trough of the South China Sea during the last 120 ka and their paleoclimatic implications. Mar. Geol. Quat. Geol. 2013, 33, 89–96, (In Chinese with English Abstract). [Google Scholar] [CrossRef]
  94. Wang, L.; Jian, Z.; Chen, J. Late Quaternary pteropods in the South China Sea: Carbonate preservation and paleoenvironmental variation. Mar. Micropaleontol. 1997, 32, 115–126. [Google Scholar] [CrossRef]
  95. Liu, Z.; Colin, C.; Trentesaux, A.; Blamart, D.; Bassinot, F.; Siani, G.; Sicre, M.-A. Erosional history of the eastern Tibetan Plateau since 190 kyr ago: Clay mineralogical and geochemical investigations from the southwestern South China Sea. Mar. Geol. 2004, 209, 1–18. [Google Scholar] [CrossRef]
  96. Chen, M.-T.; Shiau, L.-J.; Yu, P.-S.; Chiu, T.-C.; Chen, Y.-G.; Wei, K.-Y. 500,000-year records of carbonate, organic carbon, and foraminiferal sea-surface temperature from the southeastern South China Sea (near Palawan Island). Palaeogeogr. Palaeoclimatol. Palaeoecol. 2003, 197, 113–131. [Google Scholar] [CrossRef]
  97. Zhang, J.Y.; Zhao, L.; Li, B.; Li, X.J.; Zhong, H.X.; Tian, C.J. Carbonate cycle in sub-bottom strata in the South China Sea and the east sea area of Taiwan Island. Geol. China 2020, 47, 1486–1500. [Google Scholar]
  98. Li, L.; Wang, H.; Luo, B.; He, J. The characterizations and paleoceanographic significances of organic and inorganic carbon in northern South China Sea during past 40 ka. Mar. Geol. Quat. Geol. 2008, 28, 79–85, (In Chinese with English Abstract). [Google Scholar]
  99. Sun, Y.; Wu, F.; Clemens, S.C.; Oppo, D.W. Processes controlling the geochemical composition of the South China Sea sediments during the last climatic cycle. Chem. Geol. 2008, 257, 240–246. [Google Scholar] [CrossRef]
  100. Oppo, D.W.; Sun, Y. Amplitude and timing of sea-surface temperature change in the northern South China Sea: Dynamic link to the East Asian monsoon. Geology 2005, 33, 785–788. [Google Scholar] [CrossRef]
  101. Fan, W.; Chen, R.; Zhao, Q.; Gao, S. Planktonic foraminifera in core 191PC from the northern South China Sea and their paleoenvironmental significance. Acta Micropalaeontol. Sin. 2007, 24, 233–241, (In Chinese with English Abstract). [Google Scholar]
  102. Zheng, H.; Chen, G.; Xie, X.; Mei, X.; Li, J.; Ge, H.; Huang, E. Grain size distribution and dynamic control of late quaternary terrigenous sediments in the South China Sea and their implication for East Asian monsoon evolution. Quat. Sci. 2008, 28, 414–424, (In Chinese with English Abstract). [Google Scholar]
  103. Liu, Z.; Li, X.; Colin, C.; Ge, H. A high-resolution clay mineralogical record in the northern South China Sea since the Last Glacial Maximum, and its time series provenance analysis. Chin. Sci. Bull. 2010, 55, 4058–4068. [Google Scholar] [CrossRef]
  104. Xie, X.; Zheng, H.-B.; Qiao, P.-J. Millennial climate changes since MIS 3 revealed by element records in deep-sea sediments from northern South China Sea. Chin. Sci. Bull. 2014, 59, 776–784. [Google Scholar] [CrossRef]
  105. Yang, W.G.; Zheng, H.B.; Xie, X.; Zhou, B.; Cheng, X. East Asian summer monsoon maximum records in northern South China Sea during the early Holocene. Quat. Sci. 2008, 28, 425–430, (In Chinese with English Abstract). [Google Scholar]
  106. Ye, L.; Chu, F.; Ge, Q.; Xu, D. A rapid gas hydrate dissociation in the northern South China Sea since the late Younger Dryas. Earth Sci. J. China Univ. Geosci. 2013, 38, 1299–1308, (In Chinese with English Abstract). [Google Scholar]
  107. Jian, Z.; Wang, L.; Kienast, M.; Sarnthein, M.; Kuhnt, W.; Lin, H.; Wang, P. Benthic foraminiferal paleoceanography of the South China Sea over the last 40,000 years. Mar. Geol. 1999, 156, 159–186. [Google Scholar] [CrossRef]
  108. Li, J.R. Carbon Reservoir in Low-Latitude Oceans and Orbital Cycles of Monsoon Climate. Ph.D. Thesis, Tongji University, Shanghai, China, 2007. (In Chinese with English Abstract). [Google Scholar]
  109. Lehner, B.; Grill, G. Global river hydrography and network routing: Baseline data and new approaches to study the world’s large river systems. Hydrol. Process. 2013, 27, 2171–2186. [Google Scholar] [CrossRef]
  110. Li, C.; Lin, J.; Kulhanek, D. South China Sea Tectonics. In Proceedings of the International Ocean Discovery Program Expedition 349, Hong Kong, China, 26 January–30 March 2014; International Ocean Discovery Program College Station: College Station, TX, USA, 2015. [Google Scholar]
  111. Zeeden, C.; Hilgen, F.; Westerhold, T.; Lourens, L.; Röhl, U.; Bickert, T. Revised Miocene splice, astronomical tuning and calcareous plankton biochronology of ODP Site 926 between 5 and 14.4 Ma. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2013, 369, 430–451. [Google Scholar] [CrossRef]
  112. Mei, X.; Zheng, H.B.; Huang, E.Q.; Chen, G.C.; Xie, X. Characteristics of sediments in the Southern South China Sea during last 500 ka and their paleoenvironmental significancesi. Mar. Geol. Quat. Geol. 2007, 27, 77–84. [Google Scholar]
  113. Su, X.; Liu, C.; Beaufort, L.; Barbarin, N.; Jian, Z. Differences in Late Quaternary primary productivity between the western tropical Pacific and the South China Sea: Evidence from coccoliths. Deep. Sea Res. Part II Top. Stud. Oceanogr. 2015, 122, 131–141. [Google Scholar] [CrossRef]
  114. Chen, M.T.; Huang, C.Y.; Wei, K.Y. 25,000-year late Quaternary records of carbonate preservation in the South China Sea. Palaeogeogr. Palaeoclimatol. Palaeoecol. 1997, 129, 155–169. [Google Scholar] [CrossRef]
  115. Yamamoto, M.; Sai, H.; Chen, M.T.; Zhao, M. The East Asian winter monsoon variability in response to precession during the past 150,000 yr. Clim. Past 2013, 9, 2777–2788. [Google Scholar] [CrossRef]
  116. Huang, C.Y.; Wu, S.F.; Zhao, M.X.; Chen, M.T.; Wang, C.H.; Tu, X.; Yuan, P.B. Surface ocean and monsoon climate variability in the South China Sea since the last glaciation. Mar. Micropaleontol. 1997, 32, 71–94. [Google Scholar] [CrossRef]
  117. Higginson, M.J.; Maxwell, J.R.; Altabet, M.A. Nitrogen isotope and chlorin paleoproductivity records from the Northern South China Sea: Remote vs. local forcing of millennial- and orbital-scale variability. Mar. Geol. 2003, 201, 223–250. [Google Scholar] [CrossRef]
  118. Huang, C.J. The current status of cyclostratigraphy and astrochronology in the Mesozoic. Earth Sci. Front. 2014, 21, 48–66. [Google Scholar]
  119. Luz, B.; Shackleton, N.J. CaCO3 solution in the tropical East Pacific during the past 130,000 years. In Dissolution of Deep-sea Carbonates; Sliter, W.V., Bé, A.W.H., Berger, W.H., Eds.; Cushman Foundation for Foraminiferal Research: Lawrence, KS, USA, 1975; Volume 13. [Google Scholar]
  120. Anderson, R.F.; Fleisher, M.Q.; Lao, Y.; Winckler, G. Modern CaCO3 preservation in equatorial Pacific sediments in the context of late-Pleistocene glacial cycles. Mar. Chem. 2008, 111, 30–46. [Google Scholar] [CrossRef]
  121. Luo, Y.M.; Kienast, M.; Boudreau, B.P. Invariance of the carbonate chemistry of the South China Sea from the glacial period to the Holocene and its implications to the Pacific Ocean carbonate system. Earth Planet. Sci. Lett. 2018, 492, 112–120. [Google Scholar] [CrossRef]
  122. Ge, Q.; Chu, F.; Xue, Z.; Liu, J.P.; Du, Y.; Fang, Y. Paleoenvironmental records from the northern South China Sea since the Last Glacial Maximum. Acta Oceanol. Sin. 2010, 29, 46–62. [Google Scholar] [CrossRef]
  123. Yancheva, G.; Nowaczyk, N.R.; Mingram, J.; Dulski, P.; Schettler, G.; Negendank, J.F.W.; Liu, J.; Sigman, D.M.; Peterson, L.C.; Haug, G.H. Influence of the intertropical convergence zone on the East Asian monsoon. Nature 2007, 445, 74–77. [Google Scholar] [CrossRef] [PubMed]
  124. Yuan, D.; Cheng, H.; Edwards, R.L.; Dykoski, C.A.; Kelly, M.J.; Zhang, M.; Qing, J.; Lin, Y.; Wang, Y.; Wu, J.; et al. Timing, duration, and transitions of the last interglacial Asian monsoon. Science 2004, 304, 575–578. [Google Scholar] [CrossRef]
  125. Dykoski, C.A.; Edwards, R.L.; Cheng, H.; Yuan, D.; Cai, Y.; Zhang, M.; Lin, Y.; Qing, J.; An, Z.; Revenaugh, J. A high-resolution, absolute-dated Holocene and deglacial Asian monsoon record from Dongge Cave, China. Earth Planet. Sci. Lett. 2005, 233, 71–86. [Google Scholar] [CrossRef]
  126. Wen, X.S.; Zhao, H.T.; Zhang, M.Q.; Song, J.C. Sedimentary characteristics and environmental evolution of drilling core in Lingdingyang. Acta Oceanol. Sin. 1997, 19, 121–128. (In Chinese) [Google Scholar]
  127. Shao, X.; Wang, Y.; Cheng, H.; Kong, X.; Wu, J.; Edwards, R.L. Long-term trend and abrupt events of the Holocene Asian monsoon inferred from a stalagmite δ10O record from Shennongjia in Central China. Chin. Sci. Bull. 2006, 51, 221–228. [Google Scholar] [CrossRef]
  128. Alley, R.B.; Mayewski, P.A.; Sowers, T.; Stuiver, M.; Taylor, K.C.; Clark, P.U. Holocene climatic instability: A prominent, widespread event 8200 yr ago. Geology 1997, 25, 483–486. [Google Scholar] [CrossRef]
  129. Rohling, E.J.; Pälike, H. Centennial-scale climate cooling with a sudden cold event around 8,200 years ago. Nature 2005, 434, 975–979. [Google Scholar] [CrossRef]
  130. Yu, G.; Liu, J. Geological Records of Volcanic Explosions during the Last 12000 Years and the Volcanic Impacts on Climate Changes. J. Lake Sci. 2003, 15, 11–20, (In Chinese with English Abstract). [Google Scholar]
  131. Meyers, P.A. Organic geochemical proxies of paleoceanographic, paleolimnologic, and paleoclimatic processes. Org. Geochem. 1997, 27, 213–250. [Google Scholar] [CrossRef]
  132. Berger, A. Long-term variations of daily insolation and Quaternary climatic changes. J. Atmos. Sci. 1978, 35, 2362–2367. [Google Scholar] [CrossRef]
  133. Wen, X.; Liu, Z.; Wang, S.; Cheng, J.; Zhu, J. Correlation and anti-correlation of the East Asian summer and winter monsoons during the last 21,000 years. Nat. Commun. 2016, 7, 11999. [Google Scholar] [CrossRef]
  134. He, J.; Zhao, M.; Wang, P.; Li, L.; Li, Q. Changes in phytoplankton productivity and community structure in the northern South China Sea during the past 260 ka. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2013, 392, 312–323. [Google Scholar] [CrossRef]
  135. Su, X.; Liu, C.; Beaufort, L.; Tian, J.; Huang, E. Late Quaternary coccolith records in the South China Sea and East Asian monsoon dynamics. Glob. Planet. Change 2013, 111, 88–96. [Google Scholar] [CrossRef]
  136. Li, D.W.; Zhao, M.X.; Chen, M.T. East Asian winter monsoon controlling phytoplankton productivity and community structure changes in the southeastern South China Sea over the last 185 kyr. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2014, 414, 233–242. [Google Scholar] [CrossRef]
  137. Tamburini, F.; Adatte, T.; Föllmi, K.; Bernasconi, S.M.; Steinmann, P. Investigating the history of East Asian monsoon and climate during the last glacial–interglacial period (0–140,000 years): Mineralogy and geochemistry of ODP Sites 1143 and 1144, South China Sea. Mar. Geol. 2003, 201, 147–168. [Google Scholar] [CrossRef]
  138. Li, J.Y. Quaternary Diatoms from the South Chinaa Sea, Legy 184, Site1144 and Their Palaeoenvironmental Evolution. Geol. Rev. 2002, 48, 542–551, (In Chinese with English Abstract). [Google Scholar]
  139. Xue, H.; Chai, F.; Pettigrew, N.; Xu, D.; Shi, M.; Xu, J. Kuroshio intrusion and the circulation in the South China Sea. J. Geophys. Res. Ocean. 2004, 109, C02017. [Google Scholar] [CrossRef]
  140. Liu, J.; Li, T.; Xiang, R.; Chen, M.; Yan, W.; Chen, Z.; Liu, F. Influence of the Kuroshio Current intrusion on Holocene environmental transformation in the South China Sea. Holocene 2013, 23, 850–859. [Google Scholar] [CrossRef]
  141. Chen, C.C.; Jan, S.; Kuo, T.H.; Li, S.-Y. Nutrient flux and transport by the Kuroshio east of Taiwan. J. Mar. Syst. 2017, 167, 43–54. [Google Scholar] [CrossRef]
  142. Gerotto, A.; Figueira, R.C.L.; Ferreira, P.A.d.L.; Hahn, A.; Hernandez-Almeida, I.; Nagai, R.H. Carbonate-rich sequences in the deep sea of the eastern South China Sea subbasin (IODP Hole U1431D) during the last 600 kyr. Mar. Micropaleontol. 2020, 158, 101879. [Google Scholar] [CrossRef]
  143. Wang, P. Global monsoon in a geological perspective. Chin. Sci. Bull. 2009, 54, 1113–1136. [Google Scholar] [CrossRef]
Figure 2. Variations in lithology, grain size fractions, and carbonate content throughout the core. The black triangle marks the position of AMS 14C dating.
Figure 2. Variations in lithology, grain size fractions, and carbonate content throughout the core. The black triangle marks the position of AMS 14C dating.
Jmse 12 00438 g002
Figure 5. The initial age model and sedimentation rates for core SCS1. Three age control points, identified through carbonate stratigraphy, are marked with black dots and were instrumental in constructing the age model.
Figure 5. The initial age model and sedimentation rates for core SCS1. Three age control points, identified through carbonate stratigraphy, are marked with black dots and were instrumental in constructing the age model.
Jmse 12 00438 g005
Figure 6. Cyclostratigraphic interpretation of core SCS1 through various analytical lenses: (A) Displays the detrended carbonate data alongside the output from a 20 kyr cycle filter applied to the initial time series, employing a Gaussian bandpass filter of 0.05 ± 0.010 cycles/kyr. (B) Shows the evolutionary fast Fourier transform (FFT) spectrogram for the initial carbonate time series, offering insights into the frequency modulation over time. (C) Presents the 2π multi-taper method (MTM) power spectrum of the initial carbonate time series, highlighting dominant cyclical patterns. (D) Illustrates the 20 kyr cycle component extracted from precession data, utilizing a Gaussian bandpass filter of 0.05 ± 0.010 cycles/kyr for precision. (E) Depicts the tuned time domain alongside the 20 kyr filters of the detrended carbonate series, demonstrating the impact of astronomical tuning. (F) Summer insolation at 30° N [65]. (G) Features of the tuned carbonate time series. (H) Provides the evolutionary FFT spectrogram of the tuned carbonate series, showing the effect of tuning on frequency patterns. (I) Offers the 2π MTM power spectrum of the tuned carbonate series, evidencing the successful isolation of significant cycles through tuning.
Figure 6. Cyclostratigraphic interpretation of core SCS1 through various analytical lenses: (A) Displays the detrended carbonate data alongside the output from a 20 kyr cycle filter applied to the initial time series, employing a Gaussian bandpass filter of 0.05 ± 0.010 cycles/kyr. (B) Shows the evolutionary fast Fourier transform (FFT) spectrogram for the initial carbonate time series, offering insights into the frequency modulation over time. (C) Presents the 2π multi-taper method (MTM) power spectrum of the initial carbonate time series, highlighting dominant cyclical patterns. (D) Illustrates the 20 kyr cycle component extracted from precession data, utilizing a Gaussian bandpass filter of 0.05 ± 0.010 cycles/kyr for precision. (E) Depicts the tuned time domain alongside the 20 kyr filters of the detrended carbonate series, demonstrating the impact of astronomical tuning. (F) Summer insolation at 30° N [65]. (G) Features of the tuned carbonate time series. (H) Provides the evolutionary FFT spectrogram of the tuned carbonate series, showing the effect of tuning on frequency patterns. (I) Offers the 2π MTM power spectrum of the tuned carbonate series, evidencing the successful isolation of significant cycles through tuning.
Jmse 12 00438 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, C.; Zhang, W.; Zhang, C.; Zheng, L.; Yan, S.; Ma, Y.; Dang, W. Astronomical Time Scale of the Late Pleistocene in the Northern South China Sea Based on Carbonate Deposition Record. J. Mar. Sci. Eng. 2024, 12, 438. https://0-doi-org.brum.beds.ac.uk/10.3390/jmse12030438

AMA Style

Zhang C, Zhang W, Zhang C, Zheng L, Yan S, Ma Y, Dang W. Astronomical Time Scale of the Late Pleistocene in the Northern South China Sea Based on Carbonate Deposition Record. Journal of Marine Science and Engineering. 2024; 12(3):438. https://0-doi-org.brum.beds.ac.uk/10.3390/jmse12030438

Chicago/Turabian Style

Zhang, Chunhui, Wanyi Zhang, Chengjun Zhang, Liwei Zheng, Shiyi Yan, Yuanhao Ma, and Wei Dang. 2024. "Astronomical Time Scale of the Late Pleistocene in the Northern South China Sea Based on Carbonate Deposition Record" Journal of Marine Science and Engineering 12, no. 3: 438. https://0-doi-org.brum.beds.ac.uk/10.3390/jmse12030438

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop