Next Article in Journal
Laser Powder Bed Fusion of a Topology Optimized and Surface Textured Rudder Bulb with Lightweight and Drag-Reducing Design
Previous Article in Journal
A Heuristic Approach for Inter-Facility Comparison of Results from Round Robin Testing of a Floating Wind Turbine in Irregular Waves
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetic Properties and Initiation of Biogenic Reefs in Xisha Islands, South China Sea, at the Oligo–Miocene Boundary

1
State Key Laboratory of Marine Geology, Tongji University, Shanghai 200092, China
2
China National Offshore Oil Corporation Ltd., Hainan Branch, Haikou 570100, China
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2021, 9(9), 1031; https://doi.org/10.3390/jmse9091031
Submission received: 13 August 2021 / Revised: 9 September 2021 / Accepted: 15 September 2021 / Published: 19 September 2021
(This article belongs to the Section Geological Oceanography)

Abstract

:
Biogenic reefs and carbonate platforms are valuable natural resources, playing an important role in modulating the global climate and in carbon cycles through biological processes. Biogenic reefs in the Xisha (Paracel) Islands began in the late Oligocene and covaried with the deep-sea basin of the South China Sea and with the aeolian deposit in the Chinese Loess Plateau. Core XK-1 was drilled into the Xisha Islands to their granitic base and well dated by magnetostratigraphy, offering an opportunity to reveal the details of how the Xisha reefs initiated. In this report, the lower section of the biogenic reefs (23.0–24.5 Ma) was sampled for studying magnetic properties. The main results are as follows: (1) magnetic minerals in the XK-1 biogenic reefs are dominated by low-coercivity and relatively coarse-grained magnetite; (2) the variabilities of magnetic parameters can be clustered into two sections around 23.6 Ma, and the differences between the two units are evident both in the amplitudes and the means; and (3) changes in the concentration-dependent magnetic parameters can be well correlated with the records of global deep-sea oxygen and carbon isotopes, and the sea level during the Oligo–Miocene boundary. Based on these results, a close link was inferred between biogenic reef evolution in the Xisha Islands and global climate change. This link likely highlights the covariation or the dominant role of the Asian monsoon in biogenic reefs and involves different responses to global temperature, CO2, and sea-level changes on various timescales. Therefore, we proposed that the origin of biogenic reefs in the Xisha Islands was likely paced by orbital obliquity from a long-term perspective.

1. Introduction

Tropical reefs are composed of corals, algae, and other reef-building organisms with an anti-wave structure. Accounting for a quarter of the global annual carbonate production [1,2], biogenic reefs are critical in revealing the links and interactions between mid- and high-latitude environmental processes from the perspective of tropical oceans. However, they are all suffering a continuous decline and bleaching in the context of global warming and ocean acidification [3,4,5]. To overturn this decline and bleaching and to maintain ecological function again, the full cooperation of science, society, and policy in protecting coral reefs is necessary around the world [6], and due to the unclear response of coral reefs to global changes, there are many uncertainties in predicting the future of coral reefs. For example, combining human social systems with factors of forestry, freshwater lakes, fisheries, agricultural abandonment, and climate change in box models [7,8,9,10,11,12,13], an increase in complexity and in prediction uncertainty is evident [14].
Modern observations cover the past several decades, and these data have great limitations in effectively developing models for coral reefs. For example, the changing hydrochemical characteristics in geological timescales may control coral evolution [15], and during a warm period of global climate in the middle Miocene [16], coral reefs largely developed in the Western Pacific [17,18]. Hence, it is speculated that coral reefs in different geological periods have different responses to global climate change, and studying the long-term sequence of coral reefs can provide a useful reference, supplementary to modern observations.
The South China Sea is the largest marginal sea in Asia (Figure 1). Numerous coral reefs developed in the Eocene and early Oligocene in the south [19,20,21,22,23], while in the north, reefs mainly formed on uplifted fault blocks or on volcanic seamounts [17,24] starting from the late Oligocene [18,25]. These carbonate deposits not only relate to the evolution of the South China Sea but also serve as a valuable natural resource and precious heritage. Lithological studies suggest that these carbonate deposits in both the southern and northern parts of the South China Sea are comparable with each other [26,27,28], while few geochronological investigations from a paleoenvironmental and palaeoclimatological perspective have been performed (e.g., [18,29,30]), precluding our understanding of the initiation and development of these biogenic reefs. Moreover, although commonly used proxies, such as geochemical components and carbonate δ18O and δ13C, are significantly influenced by diageneses and dolomitization (e.g., [31]), the magnetic properties of biogenic reefs are useful as a paleoenvironmental proxy in the South China Sea and the South Pacific (e.g., [32,33]). Those magnetic particles are mainly magnetite in the pseudo-single domain (PSD), which could be trapped during coral growth. In addition, aeolian deposits were found on the top of biogenic reefs at several sites in the Xisha (Paracel) Islands, suggesting an evident influence of the Asian monsoon in this region [26,34,35]. Therefore, this study aims to reveal the evolution of biogenic reefs in the Xisha (Paracel) Islands and how this carbonate bank responded to global climate change during the Oligo–Miocene transition. To this end, detailed magnetic properties are investigated on core XK-1, drilled from Shidao Island in the Xisha (Paracel) Islands (Figure 1).

2. Materials and Methods

2.1. Core XK-1

The Xisha (Paracel) Islands are located in the northwest of the South China Sea (Figure 1a), and the total area of coral reefs in this region is about 1836 km2. The Xisha Uplift is the main structural element in the study area, which subsided around the Oligo–Miocene boundary [36,37]. Since then, tropical biogenic reefs have developed, forming the Xisha carbonate bank [18,30].
Core XK-1 was drilled to the bottom of the Xisha carbonate platform on Shidao Island in the Xuande Atoll, Xisha (Paracel) Islands (16°50′45″ N, 112°20′50″ E; Figure 1b), by the China National Offshore Oil Corporation in 2012–2013. The total length of the core was 1268 m, with an average recovery rate of 80%. Based on sedimentary facies, five lithological units were identified (Figure 1c): Sanya, Meishan, Huangliu, Yinggehai, and Ledong Formations (Fms.) from bottom to top [38,39]. The ages of each formation are estimated as 2.2 Ma, 5.7 Ma, 10.4 Ma, 16.6 Ma, and 24.3 Ma, respectively [18,25,32,40].
The Sanya Fm. (1258–1032 m) contains a thick dolomitic reef with some bioclastic limestone in the upper part, alternating reefal and bioclastic limestone in the middle, and brownish breccia and coral reefs in the bottom (Figure 2). Fossils of Actinacis and Spiroclypeus were found at 1245 m and 1255 m in core XK-1, respectively [41], supporting the estimate that these biogenic reefs date back to the late Oligocene [18]. Besides the two fossils, red algae, including Lithoporella melobesioides, Lithoporella minus, Neogoniolithon, and Mesophyllum, and benthic foraminifera, including Amphistegina lessonii, Amphistegina lobifera, and Spiroclypeus were also found in the lower part of the Sanya Fm. [41].
The depositional rates of the Sanya Fm. are 10–30 m/Myr, lower than other parts of the core [18]. Due to the low sampling rate (1–3 m) and the low magnetic susceptibility [18,40], there are lots of uncertainties in the astronomical age model in the lower part of core XK-1. Considering the relatively stable depositional processes of biogenic reefs in the South China Sea in the Neogene [30] and the reliability of magnetostratigraphy [25], which was supported by the identified fossils (Actinacis and Spiroclypeus), we employed the preliminary age model in this study (Figure 2c) to reveal the relationship between the initiation of biogenic reefs in the South China Sea and global climate change.

2.2. Measurements

Focusing on the interval of the Oligo–Miocene transition, the lower part of the Sanya Fm. in core XK-1 was sampled (1258–1225 m), spanning 24.5–23.0 Ma according to the preliminary age model [18]. A total of 93 samples were collected for the magnetic measurements (Figure 2).
All samples were placed in standard 8 cm3 plastic cubic boxes. Magnetic susceptibility (MS) was measured using an AGICO MFK1–FA Multi-Frequency Kappabridge magnetic susceptibility meter by [18]; however, due to low values close to the instrument background, MS data were not included in this study.
Anhysteretic remanent magnetization (ARM) was imparted onto the samples using a peak alternating field (AF) of 100 mT and a direct biasing field of 0.05 mT using a 2G Enterprises SQUID magnetometer with inline AF coils, and then, the susceptibility of ARM (χARM) was calculated as ARM/0.05 mT. Isothermal remanent magnetization (IRM) was produced with a 2-G Enterprises model 660 pulse magnetizer successively in pulsed fields of 1 T (saturated IRM, SIRM), −0.1T (IRM−0.1T), and −0.3 T (IRM−0.3T). The hard isothermal remanent magnetization (HIRM), Sratio, and Lratio, were calculated using the following equations [42]:
HIRM = SIRM IRM 0.3 T 2 ;   S ratio = IRM 0.3 T SIRM × 100 % ;   L ratio = SIRM IRM 0.3 T SIRM IRM 0.1 T
Hysteresis loop and first-order reversal curve (FORC) analyses were conducted on representative samples using a Princeton Measurements Inc. MicroMag 3900 Vibrating Sample Magnetometer (VSM). A peak field of 1 T was set for hysteresis loops, and saturation magnetization (Ms), saturation remanence (Mrs), coercive force (Bc), and the coercivity of the remanence (Bcr) were determined from the hysteresis loops [42] after they were corrected using the data between 0.7 and 1.0 T. Setting a peak field of 1.0 T and an interval of 3.2 mT, FORC diagrams (125 lines) were produced using FORCme software [43,44] with a smoothing factor of 11. These measurements were conducted at the Paleomagnetism and Geochronology Lab (PGL), Institute of Geology and Geophysics, Chinese Academy of Sciences.
Due to the extremely low content of magnetic minerals [40], high-resolution magnetic scanning was conducted on a slice (1239 m) to reveal the occurrence state of magnetic particles in biogenic reefs, at Charles University in Prague, Czech Republic [45]. Due to low remanence, the sample was first magnetized using a Magnetic Measurements Pulse Magnetizer (MMPM 10) with a field of 3 T. Then, the sample surface was scanned to reproduce the magnetic field of the sample. The sensitivity of the magnetic probe was 0.01 μT, the distance between the probe and the sample was <0.1 mm, and the spatial resolution was 200 μm.

3. Results

3.1. Changes in Magnetic Parameters

χARM is usually sensitive to single-domain (SD) grains [46,47], and SIRM can be used to infer magnetic particles excluding the influence of superparamagnetic (SP) grains [48]. The values of χARM and SIRM are 0.16–9.35 × 10−8 m3/kg and 6.0–1310 × 10−3 Am2/kg, respectively (Figure 3), inferring a low content of magnetic minerals in biogenic reefs.
Low Lratio values with a high standard deviation (3.96 ± 3.33), close to those in deep-sea sediments in the western Pacific [49], indicate the dominance of low-coercivity magnetic minerals, while low Sratio values (0.61 ± 0.17) might infer the contribution of high-coercivity minerals in biogenic reefs (Figure 3). However, a low concentration of magnetic minerals and a high CaCO3 content may introduce large uncertainties into magnetic measurements. For example, the relationship between the magnetic parameters are more discrete (Figure 4) than those in other sediments (e.g., [49,50,51]). Considering that the concentration-dependent magnetic parameters (χARM and IRMs) have similar variations throughout the studied section (Figure 3), it is inferred that the discrete relationship between the magnetic parameters (Figure 4) may introduce large uncertainties into calculations. Therefore the difference between Lratio and Sratio does not indicate the dominance of high-coercivity magnetic minerals in biogenic reefs of core XK-1.
These magnetic variabilities can be generally divided into two units (Figure 3). In specific, for the upper part, the concentration-dependent magnetic parameters (χARM and IRMs) have a large average, while for the low part, these records vary greatly but with lower values. By a one-way analysis of variance (ANOVA), the difference of all parameters (χARM and IRMs) between the two units is statistically significant (p < 0.05), confirming a shift of the changes in magnetic parameters around 23.6 Ma. In addition, the difference in the relationships is not evident (Figure 4), inferring that the source and/or properties of magnetic minerals in biogenic reefs in core XK-1 are similar across the Oligo–Miocene transition.

3.2. Magnetic Minerals and the Proxy

Rock magnetic analysis shows that there are high levels of noise in the hysteresis loops for most of the samples (Figure 5a), and only one obtains a smooth curve (Figure 5b), demonstrating the significant influence of low magnetic minerals and high CaCO3 content. For the smoothed loop (Figure 5b), it is closed above 200 mT, with Bc and Bcr of 11 mT and 52 mT, respectively. On the Day plot, these data are distributed within the lower part of the PSD range. The large vertical spread and wide horizonal spread of the contours in the FORC diagram (Figure 5c) indicate that magnetostatic interactions are evident [52,53], and the bulk of the coercivity distribution lies in the 5–40 mT range, with peaks at ~10 mT. All of these magnetic properties agree well with those in the upper part of core XK-1 [40], possibly inferring the dominance of detrital magnetite in biogenic reefs.
Although detrital minerals are not evident in slice samples under a microscope and most of them were identified by SEM images and energy spectrum analysis (e.g., [54]), high-resolution magnetic scanning can provide useful information about the distribution of magnetic minerals in biogenic reefs [40]. As shown (Figure 5d), the magnetic field of the sample surface is generally weak in the middle part and relatively high at the lower right corner. It is suggested that the content of magnetic minerals in biogenic reefs can be used to indicate the growth intensity [18,30], since magnetic minerals may be enclosed during growth. The relationship is as follows: the greater the magnetic parameters, the higher the concentration of magnetic minerals and the stronger the growing (deposition) processes of biogenic reefs. This case is similarly reported in the Tahiti coral reef in the Southern Pacific, and magnetite particles were locked in biogenic reefs by biological (microbial) processes [33].
Based on these magnetic properties and the high similarity between the concentration-dependent magnetic parameters (χARM and IRMs), a magnetic proxy (MagXK) indicating the growing process of biogenic reefs in core XK-1 was derived (Figure 6a) by a principle component analysis on χARM and IRMs. As a result, the MagXK inherits a similar pattern of four magnetic parameters, accounting for 87.8% of variance in total (Table 1), and varies with some cycles (Figure 6b). For other components with a much lower eigenvalue (Table 1), they are not included in analysis. By a wavelet analysis, these cycles can be identified as 40–50 kyr, 100–200 kyr, and 400–500 kyr, which are generally correlated to the periodicities of Earth’s orbits and discussed below.

4. Discussion

Biogenic reefs account for a quarter of global annual carbonate production [1,2]. However, these reefs are suffering a continuous decline and bleaching due to global warming and ocean acidification [3,4,5]. Five principals are critical in the development of carbonate platforms (e.g., [17,57,58,59,60]): carbonate production, tectonic activities, sea-level changes, terrigenous material inputs, and paleoceanography. Considering that each principal usually covaries, unmixing the MagXK record of biogenic reefs on various timescales offers an opportunity to test the roles of each factor (Figure 7 and Figure 8).
As shown in a comparison between orbital eccentricity and the MagXK at the ~100 kyr band (Figure 6c), there is a transition around the Oligo–Miocene boundary. Additionally, in the early Miocene, the two records covaried at the 405 kyr band (Figure 6a). The changing relationship may infer a complex response of biogenic reefs at the very beginning to various environmental processes, and this is made evident by comparing this instance to other paleoenvironmental proxies (Figure 7).
For example, using δ18O and δ13C to indicate global temperature and CO2 changes, respectively [16,63], the two key processes are closely correlated (Figure 7a,b). As a response to global climate change, on the ~0.5 Myr timescale, the MagXK record is positively correlated with global temperature changes (δ18O), indicating that biogenic reefs tend to decay in a cooler climate, while on the 0.5 Myr timescale, the MagXK is in-phase linked to global CO2 changes (δ13C), suggesting that reef flourishing is in a higher CO2 state. Similarly, the MagXK was positively correlated with the global sea level on the ~0.5 Myr timescale, suggesting that high sea-levels would not prevent reef developments, while on the 0.5 Myr timescale, it showed the opposite relationship (Figure 7d). Moreover, on the ~100–200 kyr timescales, the MagXK is generally correlated with global temperature and CO2 changes (δ18O and δ13C) (Figure 7e,f) as well as global sea level with some anti-phase intervals (Figure 7h).
Based on these observations, we present a complex relationship for the initiation of biogenic reefs in the South China Sea at the Oligo–Miocene boundary. In specific, on high-frequency variabilities, reef development in the South China Sea covaried with global temperature and CO2 changes (δ18O and δ13C) and sea levels. On a medium timescale, high temperature (δ18O), high sea level, and low CO213C) would induce biogenic reefs, and for long-term changes on the 0.5 Myr timescale, biogenic reefs were closely linked to low temperature (δ18O), low sea level, and high CO213C).
In modern studies, ocean acidification due to atmospheric CO2 can significantly affect coral calcification rates, inhibit the development of coral larvae, and trigger the dissolution of coral reefs [64,65,66,67,68], and global warming with high sea levels can induce coral bleaching and inhibit self-renovation of coral reefs [4,6,14,69,70,71]. From a long-term perspective, our results may suggest a nonlinear and more complex response of reef development to global climate change, and this is also evident in modern observation. For example, corals can resist heat stress to a certain extent by changing the types of symbiotic algae and by regulating gene expression [72,73,74].
On the other hand, there are some offsets in these comparisons (Figure 7), likely suggesting that factors besides temperature, CO2, and sea level changes played an important role in the initiation of biogenic reefs in the South China Sea. For example, based on dating carbonate in the Maldives, the South Asian monsoon was proposed to weakly begin at ~25 Ma with an abrupt onset at ~13 Ma [75]. Meanwhile, the East Asian monsoon similarly initiated in the late Oligocene [61,76]. In a comparison of the MagXK to monsoon changes inferred from the loess magnetic susceptibility of the Zhuanglang Profile in the Chinese Loess Plateau (CLP-ZL MS), a better but distinct agreement was observed in various timescales (Figure 7c,g). In specific, on the 0.5 Myr timescale, the MagXK is negatively correlated with the CLP-ZL MS record, while on the ~100–200 kyr timescales, a positive relationship is evident, confirming the reliability of the age-depth model within the study interval. Due to the anti-phase relationship between the East Asian summer and winter monsoons (e.g., [61,77]), it is inferred that the winter monsoon may be the dominant factor on the 0.5 Myr timescale while the summer monsoon may dominate on the ~100–200 kyr timescales.
In the Xisha (Paracel) Islands, the sea surface temperature is above 29 °C from May to September and below 24 °C in January, and thus, winter (from October to December) is the more favorable season for reef development [78]. Integrating all of this evidence, the link between biogenic reefs in the South China Sea and global climate change during 24.5–23.0 Ma was proposed as follows.
At the Oligo–Miocene boundary, subject to regional turbidity, temperature, and nutrients, coral reefs in the Caribbean, Central America, were extinct [79]. Meanwhile, due to tectonic subsidence, biogenic reefs initiated in the South China Sea [18,30] and in the Indian Ocean [75]. At the very beginning of biogenic reefs in the South China Sea, the growth intensity may be closely correlated to, or paced by, the East Asian winter monsoon on the 0.5 Myr timescales, and the reefs were likely linked to the East Asian summer monsoon on the ~100–200 kyr timescales. Since the reefs migrated seaward as sea levels fell (e.g., [80]), the positive relationship demonstrates that sea level is another important factor in reef development in the study area. However, there are many offsets between the MagXK and global sea level on the ~100–200 kyr timescales, implying the difference between global and regional sea-level changes, or that the influence of sea level may not have been consistent during the depositional period. For example, during 24.0–23.6 Ma and 23.2–23.0 Ma, high growth intensity indicated by the MagXK record was associated with sea-level low stands (Figure 7h). The anti-phase intervals between the MagXK and global sea level may highlight the influence of temperature and atmospheric CO2. For global temperature and CO2 changes, distinct influences were observed on various timescales, likely illustrating a self-adjustment process in the coral system. The relationship can be summarized: (1) high temperature and low CO2 could induce reef developments on the ~0.5 Myr and ~100–200 kyr timescales, (2) low temperature with high CO2 would pace biogenic reefs on the 0.5 Myr timescales, and the latter was similarly reported in tropical reefs in the last interglacial period [81]. It is noted that the correlations presented herein are based on the preliminary age model (Figure 2c), and high-frequency variabilities are usually more sensitive to the reliability of the age model. Thus, it is possible that the correlations on the ~100–200 kyr timescales could be modified in light of future research.
For long-term variabilities, orbital parameters have been proposed to be the major force on various paleoenvironmental processes on the Earth (e.g., [16,82,83]), and orbital obliquity is thought to be more important than previously assumed (e.g., [84,85]). In this study, on the 1.0 Myr timescales, the MagXK, the CLP-ZL MS, and orbital obliquity agree well with each other, while global temperature, CO2, and sea level are coupled (Figure 8). This high similarity and in-phase variabilities between the MagXK, the CLP-ZL MS, and orbital obliquity may highlight the obliquity in pacing the initiation of the Asian monsoon and biogenic reefs in the South China Sea as the first-order force.

5. Conclusions

By studying the magnetic properties of biogenic reefs in core XK-1, drilled from the Xisha (Paracel) Islands, South China Sea, the detailed response of biogenic reefs to global climate change was revealed during the Oligo–Miocene transition. The main results are as follows: (1) magnetic minerals in the XK-1 biogenic reefs are dominated by low-coercivity and relatively coarse magnetite; (2) changes in magnetic parameters can be clustered into two sections at ~23.6 Ma, and the differences between the two units are evident. By stacking the concentration-dependent magnetic parameters (χARM and IRMs), a magnetic record indicating the development of biogenic reefs was derived. Based on this record, a complex relationship of biogenic reefs in response to global δ18O, δ13C, and sea level changes was observed. Moreover, biogenic reefs were also closely correlated or paced by the East Asian winter monsoon on the 0.5 Myr timescales. Therefore, there is a close link between biogenic reef evolution in the Xisha Islands and global climate change during the Oligo–Miocene boundary, and the origin was likely paced by orbital obliquity on long-term timescales. However, the age model of core XK-1 was only based on magnetostratigraphy and the relationship between reef growth and magnetic properties has not been supported by modern observations. Investigation about these two keys are worthy of further studies in the future.

Author Contributions

Conceptualization and methodology, L.Y.; sample collection, X.L.; formal analysis, Y.L. and W.C.; original draft preparation, X.L. and L.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key R&D Program of China (2018YFE0202401), the National Natural Science Foundation of China (42177422), and the Global Changing and Air-sea Interaction (GASI-GEOGE-04).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All magnetic data presented in this work can be obtained from L.Y. (yiliang@tongji.edu.cn) and have been stored in an online repository (https://mda.vliz.be/directlink.php?fid=VLIZ_00000824_613746d253ee3437877792) ‘Marine Data Archive’.

Acknowledgments

We thank Chenglong Deng at the Institute of Geology and Geophysics, Chinese Academy of Sciences, for help with this study and Xiaoke Qiang at the Institute of Earth Environment, Chinese Academy of Sciences, for sharing the loess data of the Zhuanglang Profile.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Milliman, J.D. Production and accumulation of calcium carbonate in the ocean: Budget of a nonsteady state. Glob. Biogeochem. Cycles 1993, 7, 927–957. [Google Scholar] [CrossRef]
  2. Milliman, J.D.; Droxler, A.W. Neritic and pelagic carbonate sedimentation in the marine environment: Ignorance is not bliss. Geol. Rundsch. 1996, 85, 496–504. [Google Scholar] [CrossRef]
  3. Eyre, B.D.; Andersson, A.J.; Cyronak, T. Benthic coral reef calcium carbonate dissolution in an acidifying ocean. Nat. Clim. Chang. 2014, 4, 969–976. [Google Scholar] [CrossRef] [Green Version]
  4. Hoegh-Guldberg, O.; Mumby, P.J.; Hooten, A.J.; Steneck, R.S.; Greenfield, P.; Gomez, E.; Harvell, C.D.; Sale, P.F.; Edwards, A.J.; Caldeira, K.; et al. Coral reefs under rapid climate change and ocean acidification. Science 2007, 318, 1737–1742. [Google Scholar] [CrossRef] [Green Version]
  5. Hughes, T.P.; Kerry, J.T.; Álvarez-Noriega, M.; Álvarez-Romero, J.G.; Anderson, K.D.; Baird, A.H.; Babcock, R.C.; Beger, M.; Bellwood, D.R.; Berkelmans, R.; et al. Global warming and recurrent mass bleaching of corals. Nature 2017, 543, 373–377. [Google Scholar] [CrossRef] [PubMed]
  6. Pandolfi, J. Incorporating Uncertainty in Predicting the Future Response of Coral Reefs to Climate Change. Annu. Rev. Ecol. Evol. Syst. 2015, 46, 281–303. [Google Scholar] [CrossRef]
  7. Cumming, G.S.; Morrison, T.H.; Hughes, T.P. New Directions for Understanding the Spatial Resilience of Social–Ecological Systems. Ecosystems 2017, 20, 649–664. [Google Scholar] [CrossRef]
  8. Levin, S.; Xepapadeas, T.; Crépin, A.-S.; Norberg, J.; de Zeeuw, A.; Folke, C.; Hughes, T.; Arrow, K.; Barrett, S.; Daily, G.; et al. Social-ecological systems as complex adaptive systems: Modeling and policy implications. Environ. Dev. Econ. 2012, 18, 111–132. [Google Scholar] [CrossRef] [Green Version]
  9. Fischer, J.; Gardner, T.A.; Bennett, E.M.; Balvanera, P.; Biggs, R.; Carpenter, S.; Daw, T.; Folke, C.; Hill, R.; Hughes, T.P.; et al. Advancing sustainability through mainstreaming a social–ecological systems perspective. Curr. Opin. Environ. Sustain. 2015, 14, 144–149. [Google Scholar] [CrossRef]
  10. Martin, R.; Schlüter, M. Combining system dynamics and agent-based modeling to analyze social-ecological interactions—an example from modeling restoration of a shallow lake. Front. Environ. Sci. 2015, 3, 66. [Google Scholar] [CrossRef] [Green Version]
  11. Laborde, S.; Fernández, A.; Phang, S.C.; Hamilton, I.M.; Henry, N.; Jung, H.C.; Mahamat, A.; Ahmadou, M.; Labara, B.K.; Kari, S.; et al. Social-ecological feedbacks lead to unsustainable lock-in in an inland fishery. Glob. Environ. Chang. 2016, 41, 13–25. [Google Scholar] [CrossRef] [Green Version]
  12. Figueiredo, J.; Pereira, H.M. Regime shifts in a socio-ecological model of farmland abandonment. Landsc. Ecol. 2011, 26, 737–749. [Google Scholar] [CrossRef]
  13. Ban, S.S.; Graham, N.A.J.; Connolly, S.R. Evidence for multiple stressor interactions and effects on coral reefs. Glob. Chang. Biol. 2014, 20, 681–697. [Google Scholar] [CrossRef] [PubMed]
  14. Hughes, T.P.; Barnes, M.L.; Bellwood, D.R.; Cinner, J.E.; Cumming, G.S.; Jackson, J.B.C.; Kleypas, J.; van de Leemput, I.A.; Lough, J.M.; Morrison, T.H.; et al. Coral reefs in the Anthropocene. Nature 2017, 546, 82–90. [Google Scholar] [CrossRef] [PubMed]
  15. Quattrini, A.M.; Rodríguez, E.; Faircloth, B.C.; Cowman, P.F.; Brugler, M.R.; Farfan, G.A.; Hellberg, M.E.; Kitahara, M.V.; Morrison, C.L.; Paz-García, D.A.; et al. Palaeoclimate ocean conditions shaped the evolution of corals and their skeletons through deep time. Nat. Ecol. Evol. 2020, 4, 1531–1538. [Google Scholar] [CrossRef] [PubMed]
  16. Zachos, J.; Pagani, M.; Sloan, L.; Thomas, E.; Billups, K. Trends, Rhythms, and Aberrations in Global Climate 65 Ma to Present. Science 2001, 292, 686–693. [Google Scholar] [CrossRef] [PubMed]
  17. Wu, S.; Yang, Z.; Wang, D.; Lü, F.; Lüdmann, T.; Fulthorpe, C.; Wang, B. Architecture, development and geological control of the Xisha carbonate platforms, northwestern South China Sea. Mar. Geol. 2014, 350, 71–83. [Google Scholar] [CrossRef]
  18. Yi, L.; Jian, Z.; Liu, X.; Zhu, Y.; Zhang, D.; Wang, Z.; Deng, C. Astronomical tuning and magnetostratigraphy of Neogene biogenic reefs in Xisha Islands, South China Sea. Sci. Bull. 2018, 63, 564–573. [Google Scholar] [CrossRef] [Green Version]
  19. Fulthorpe, C.S.; Schlanger, S.O. Paleo-oceanographic and tectonic setting of Early Miocene reefs and associated carbonates offshore Southeast Asia. AAPG Bull. 1989, 73, 729–756. [Google Scholar]
  20. Sales, A.O.; Jacobsen, E.C.; Morado Jr, A.A.; Benavidez, J.J.; Navarro, F.A.; Lim, A.E. The petroleum potential of deep-water northwest Palawan Block GSEC 66. J. Asian Earth Sci. 1997, 15, 217–240. [Google Scholar] [CrossRef]
  21. Williams, H.H. Play concepts-northwest Palawan, Philippines. J. Asian Earth Sci. 1997, 15, 251–273. [Google Scholar] [CrossRef]
  22. Hutchison, C.S. Marginal basin evolution: The southern South China Sea. Mar. Pet. Geol. 2004, 21, 1129–1148. [Google Scholar] [CrossRef]
  23. Hutchison, C.S.; Vijayan, V.R. What are the Spratly Islands? J. Asian Earth Sci. 2010, 39, 371–385. [Google Scholar] [CrossRef]
  24. Ma, Y.; Wu, S.; Lv, F.; Dong, D.; Sun, Q.; Lu, Y.; Gu, M. Seismic characteristics and development of the Xisha carbonate platforms, northern margin of the South China Sea. J. Asian Earth Sci. 2011, 40, 770–783. [Google Scholar] [CrossRef]
  25. Yi, L.; Wang, Z.; Zhang, D.; Liu, X.; You, L.; Luo, W.; Zhu, Y.; Qin, H.; Deng, C. Magnetostratigraphy of biogenic reefs in the Sanya Formation of Borehole XK-1 from Xisha Islands and its environmental significance. Coast. Eng. 2016, 35, 1–11. (In Chinese) [Google Scholar]
  26. Zhang, M.S.; He, Q.X.; Ye, Z.J. The Geologic Research of Deposition of Bioherm Carbonate in the Xisha Islands; Science Press: Beijing, China, 1989. [Google Scholar]
  27. Zhao, H.T.; Sha, Q.A.; Zhu, Y.Z. Quaternary Coral Reef: Geology of Yongshu Reef, Nansha Islands; China Ocean Press: Beijing, China, 1992; p. 264. [Google Scholar]
  28. Zhu, Y.Z.; Sha, Q.A.; Guo, L.F.; Yu, K.F.; Zhao, H.T. Cenozoic Coral Reef Geology of Yongshu Reef, Nansha Islands; Science Press: Beijing, China, 1997; p. 137. [Google Scholar]
  29. Xu, S.; Yu, K.; Fan, T.; Jiang, W.; Wang, R.; Zhang, Y.; Yue, Y.; Wang, S. Coral reef carbonate δ13C records from the northern South China Sea: A useful proxy for seawater δ13C and the carbon cycle over the past 1.8 Ma. Glob. Planet. Chang. 2019, 182, 103003. [Google Scholar] [CrossRef]
  30. Yi, L.; Deng, C.; Yan, W.; Wu, H.; Zhang, C.; Xu, W.; Su, X.; He, H.; Guo, Z. Neogene–quaternary magnetostratigraphy of the biogenic reef sequence of core NK–1 in Nansha Qundao, South China Sea. Sci. Bull. 2021, 66, 200–203. [Google Scholar] [CrossRef]
  31. Qiao, P.; Zhu, W.; Shao, L.; Zhang, D.; Cheng, X.; Song, Y. Carbonate stable isotope stratigraphy of Well Xike-1, Xisha Islands. J. China Univ. Geosci. 2015, 40, 725–732. [Google Scholar] [CrossRef]
  32. Wang, Z.; Zhang, D.; Liu, X.; You, L.; Luo, W.; Yi, L.; Zhu, Y.; Qin, H.; Xie, Q.; Che, Z.; et al. Preliminary results of rock magnetism and magnetostratigraphy for Late Miocene to Pliocene biogenic reefs in the Xisha Islands, South China Sea. Chin. J. Geophys. 2016, 59, 4178–4187. [Google Scholar] [CrossRef]
  33. Lund, S.; Platzman, E.; Thouveny, N.; Camoin, G.; Corsetti, F.; Berelson, W. Biological control of paleomagnetic remanence acquisition in carbonate framework rocks of the Tahiti coral reef. Earth Planet. Sci. Lett. 2010, 298, 14–22. [Google Scholar] [CrossRef]
  34. Wei, X.; Deng, J.F.; Xie, W.Y.; Zhu, Y.J.; Zhao, G.C.; Li, Y.X.; Chen, Y.H. Constraints on biogenetic reef formation during evolution of the South China Sea and exploration potential analysis. Earth Geosci. Front. 2005, 12, 245–252. [Google Scholar]
  35. Wang, P.; Li, Q. The South China Sea: Paleoceanography and Sedimentology; Springer: Dordrecht, The Netherlands, 2009; Volume 13, p. 506. [Google Scholar]
  36. Xu, G.Q.; Lv, B.Q.; Wang, H.G. Drown event research: Insight from Cenozoic carbonate platform in northern South China Sea. J. Tongji Univ. 2002, 30, 35–40. [Google Scholar]
  37. Fyhn, M.B.W.; Boldreel, L.O.; Nielsen, L.H.; Giang, T.C.; Nga, L.H.; Hong, N.T.M.; Nguyen, N.D.; Abatzis, I. Carbonate platform growth and demise offshore Central Vietnam: Effects of Early Miocene transgression and subsequent onshore uplift. J. Asian Earth Sci. 2013, 76, 152–168. [Google Scholar] [CrossRef]
  38. Zhu, W.; Wang, Z.; Mi, L.; Du, X.; Xie, X.; Lu, Y.; Zhang, D.; Sun, Z.; Liu, X.; You, L. Sequence stratigraphic framework and reef growth unit of Well Xike-1 from Xisha Islands, South China Sea. J. China Univ. Geosicnes 2015, 40, 677–687. [Google Scholar]
  39. Luo, W.; Zhang, D.; Liu, X.; Wang, Z.; Hu, W.; Wang, Y. A comprehensive stratigraphic study of Well XK-1 in the Xisha Area. J. Stratigr. 2018, 42, 485–498. [Google Scholar]
  40. Wang, Z.; Zhang, D.; Liu, X.; You, L.; Luo, W.; Yi, L.; Tan, L.; Zhu, Y.; Qin, H.; Cheng, H.; et al. Magnetostratigraphy and 230Th dating of Pleistocene biogenic reefs in XK-1 borehole from Xisha Islands, South China Sea. Chin. J. Geophys. 2017, 60, 1027–1038. (In Chinese) [Google Scholar] [CrossRef]
  41. Wu, F.; Xie, X.; Zhu, Y.; Chen, B.; Shang, Z. Sequence stratigraphy of the Late Oligocene carbonate system on the Xisha Islands in the South China Sea. Int. J. Earth Sci. 2021, 110, 1611–1629. [Google Scholar] [CrossRef]
  42. Tauxe, L. Essentials of Paleomagnetism; University of California Press: Berkeley, CA, USA, 2010. [Google Scholar]
  43. Heslop, D. Numerical strategies for magnetic mineral unmixing. Earth-Sci. Rev. 2015, 150, 256–284. [Google Scholar] [CrossRef]
  44. Heslop, D.; Roberts, A.P. Unmixing Magnetic Hysteresis Loops. J. Geophys. Res. Atmos. 2012, 117, 3758. [Google Scholar] [CrossRef]
  45. Kletetschka, G.; Schnabl, P.; Šifnerová, K.; Tasáryová, Z.; Manda, Š.; Pruner, P. Magnetic scanning and interpretation of paleomagnetic data from Prague Synform’s volcanics. Studia Geophys. Geod. 2013, 57, 103–117. [Google Scholar] [CrossRef]
  46. Duan, Z.; Gao, X.; Liu, Q. Anhysteretic remanent magnetization (ARM) and its application to geoscience. Prog. Geophys. 2012, 27, 1929–1938. [Google Scholar] [CrossRef]
  47. Maher, B.A. Magnetic properties of some synthetic sub-micron magnetites. Geophys. J. 1988, 94, 83–96. [Google Scholar] [CrossRef]
  48. Evans, M.E.; Heller, F. Environmental Magnetism: Principles and Applications of Enviromagnetics; Academic Press: New York, NY, USA, 2003; p. 322. [Google Scholar]
  49. Yi, L.; Wang, H.; Liu, G.; Chen, Y.; Yao, H.; Deng, X. Magnetic minerals in Mid-Pleistocene sediments on the Caiwei Guyot, Northwest Pacific and their response to the Mid-Brunhes climate event. Acta Oceanol. Sin. 2021, 40, 253. [Google Scholar]
  50. Yang, D.; Wang, M.; Lu, H.; Ding, Z.; Liu, J.; Yan, C. Magnetic properties and correlation with heavy metals in mangrove sediments, the case study on the coast of Fujian, China. Mar. Pollut. Bull. 2019, 146, 865–873. [Google Scholar] [CrossRef] [PubMed]
  51. Zhang, W.; Appel, E.; Fang, X.; Song, C.; Cirpka, O. Magnetostratigraphy of deep drilling core SG-1 in the western Qaidam Basin (NE Tibetan Plateau) and its tectonic implications. Quat. Res. 2012, 78, 139–148. [Google Scholar] [CrossRef]
  52. Roberts, A.P.; Heslop, D.; Zhao, X.; Pike, C.R. Understanding fine magnetic particle systems through use of first-order reversal curve diagrams. Rev. Geophys. 2014, 52, 557–602. [Google Scholar] [CrossRef]
  53. Roberts, A.P.; Pike, C.R.; Verosub, K.L. First-order reversal curve diagrams: A new tool for characterizing the magnetic properties of natural samples. J. Geophys. Res. Solid Earth 2000, 105, 28461–28475. [Google Scholar] [CrossRef]
  54. Wang, X.; Wang, Y.; Liu, J.; Zhang, D.; Li, X.; Shi, Z. Characteristics and significance of manganese minerals in dolostones from Well Xike-1, Xisha Area, South China Sea. Mineral. Petrol. 2020, 40, 81–91. [Google Scholar]
  55. Laskar, J.; Fienga, A.; Gastineau, M.; Manche, H. La2010: A new orbital solution for the long-term motion of the Earth. Astron. Astrophys. 2011, 532, A89. [Google Scholar] [CrossRef]
  56. Grinsted, A.; Moore, J.C.; Jevrejeva, S. Application of the cross wavelet transform and wavelet coherence to geophysical time series. Nonlinear Process. Geophys. 2004, 11, 561–566. [Google Scholar] [CrossRef]
  57. Bachtel, S.L.; Kissling, R.D.; Martono, D.; Rahardjanto, S.P.; Dunn, P.A.; MacDonald, B.A. Seismic Stratigraphic Evolution of the Miocene–Pliocene Segitiga Platform, East Natuna Sea, Indonesia: The Origin, Growth Ad Demise of an Isolated Carbonate Platform. In Seismic Imaging of Carbonate Reservoir and Systems; Eberli, G.P., Massaferro, J.L., Sarg, J.F., Eds.; AAPG Memoir: Tulsa, OK, USA, 2003; Volume 81, pp. 309–328. [Google Scholar]
  58. Belopolsky, A.V.; Droxler, A.W. Imaging Tertiary carbonate systems, the Maldives, Indian Ocean: Insights into carbonate sequence interpretation. Lead. Edge 2003, 22, 646–652. [Google Scholar] [CrossRef]
  59. Epting, M. The Miocene carbonate buildups of central Luconia, offshore Sarawak. In Atlas of Seismic Stratigraphy; Bally, A.W., Ed.; AAPG: Tulsa, OK, USA, 1989; pp. 168–173. [Google Scholar]
  60. Wilson, M.E.J. Cenozoic carbonates in Southeast Asia: Implications for equatorial carbonate development. Sediment. Geol. 2002, 147, 295–428. [Google Scholar] [CrossRef]
  61. Qiang, X.; An, Z.; Song, Y.; Chang, H.; Sun, Y.; Liu, W.; Ao, H.; Dong, J.; Fu, C.; Wu, F.; et al. New eolian red clay sequence on the western Chinese Loess Plateau linked to onset of Asian desertification about 25 Ma ago. Sci. China Earth Sci. 2011, 54, 136–144. [Google Scholar] [CrossRef]
  62. Miller, K.G.; Browning, J.V.; Schmelz, W.J.; Kopp, R.E.; Mountain, G.S.; Wright, J.D. Cenozoic sea-level and cryospheric evolution from deep-sea geochemical and continental margin records. Sci. Adv. 2020, 6, eaaz1346. [Google Scholar] [CrossRef] [PubMed]
  63. Lisiecki, L.E. A benthic δ13C-based proxy for atmospheric pCO2 over the last 1.5 Myr. Geophys. Res. Lett. 2010, 37, L21708. [Google Scholar] [CrossRef] [Green Version]
  64. Crook, E.D.; Cohen, A.L.; Rebolledo-Vieyra, M.; Hernandez, L.; Paytan, A. Reduced calcification and lack of acclimatization by coral colonies growing in areas of persistent natural acidification. Proc. Natl. Acad. Sci. USA 2013, 110, 11044–11049. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Fantazzini, P.; Mengoli, S.; Pasquini, L.; Bortolotti, V.; Brizi, L.; Mariani, M.; Di Giosia, M.; Fermani, S.; Capaccioni, B.; Caroselli, E.; et al. Gains and losses of coral skeletal porosity changes with ocean acidification acclimation. Nat. Commun. 2015, 6, 7785. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Mollica, N.R.; Guo, W.; Cohen, A.L.; Huang, K.F.; Foster, G.L.; Donald, H.K.; Solow, A.R. Ocean acidification affects coral growth by reducing skeletal density. Proc. Natl. Acad. Sci. USA 2018, 115, 1754–1759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Tambutté, E.; Venn, A.A.; Holcomb, M.; Segonds, N.; Techer, N.; Zoccola, D.; Allemand, D.; Tambutté, S. Morphological plasticity of the coral skeleton under CO2-driven seawater acidification. Nat. Commun. 2015, 6, 7368. [Google Scholar] [CrossRef] [Green Version]
  68. Foster, T.; Falter, J.; McCulloch, M.; Clode, P. Ocean acidification causes structural deformities in juvenile coral skeletons. Sci. Adv. 2016, 2, e1501130. [Google Scholar] [CrossRef] [Green Version]
  69. Frölicher, T.L.; Fischer, E.M.; Gruber, N. Marine heatwaves under global warming. Nature 2018, 560, 360–364. [Google Scholar] [CrossRef] [PubMed]
  70. Smale, D.A.; Wernberg, T.; Oliver, E.C.J.; Thomsen, M.; Harvey, B.P.; Straub, S.C.; Burrows, M.T.; Alexander, L.V.; Benthuysen, J.A.; Donat, M.G.; et al. Marine heatwaves threaten global biodiversity and the provision of ecosystem services. Nat. Clim. Chang. 2019, 9, 306–312. [Google Scholar] [CrossRef] [Green Version]
  71. Hughes, T.P.; Anderson, K.D.; Connolly, S.R.; Heron, S.F.; Kerry, J.T.; Lough, J.M.; Baird, A.H.; Baum, J.K.; Berumen, M.L.; Bridge, T.C.; et al. Spatial and temporal patterns of mass bleaching of corals in the Anthropocene. Science 2018, 359, 80–83. [Google Scholar] [CrossRef] [Green Version]
  72. Enochs, I.C.; Manzello, D.P.; Donham, E.M.; Kolodziej, G.; Okano, R.; Johnston, L.; Young, C.; Iguel, J.; Edwards, C.B.; Fox, M.D.; et al. Shift from coral to macroalgae dominance on a volcanically acidified reef. Nat. Clim. Chang. 2015, 5, 1083–1088. [Google Scholar] [CrossRef]
  73. Albright, R.; Caldeira, L.; Hosfelt, J.; Kwiatkowski, L.; Maclaren, J.K.; Mason, B.M.; Nebuchina, Y.; Ninokawa, A.; Pongratz, J.; Ricke, K.L.; et al. Reversal of ocean acidification enhances net coral reef calcification. Nature 2016, 531, 362–365. [Google Scholar] [CrossRef] [PubMed]
  74. Cooper, T.; O’Leary, R.; Lough, J. Growth of Western Australian Corals in the Anthropocene. Science 2012, 335, 593–596. [Google Scholar] [CrossRef] [Green Version]
  75. Betzler, C.; Eberli, G.P.; Kroon, D.; Wright, J.D.; Swart, P.K.; Nath, B.N.; Alvarez-Zarikian, C.A.; Alonso-García, M.; Bialik, O.M.; Blättler, C.L.; et al. The abrupt onset of the modern South Asian Monsoon winds. Sci. Rep. 2016, 6, 29838. [Google Scholar] [CrossRef] [Green Version]
  76. Guo, Z.; Ruddiman, W.F.; Hao, Q.; Wu, H.; Qiao, Y.; Zhu, R.; Peng, S.; Wei, J.J.; Yuan, B.; Liu, T.S. Onset of Asian desertification by 22 Myr ago inferred from loess deposits in China. Nature 2002, 416, 159–163. [Google Scholar] [CrossRef] [PubMed]
  77. 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]
  78. Huang, H.; Dong, Z.; Lian, J. Establishment of nature reserve of coral reef ecosystem on the Xisha Islands. Trop. Geogr. 2008, 28, 540–544. [Google Scholar]
  79. Edinger, E.N.; Risk, M.J. Oligocene-Miocene extinction and geographic restriction of Caribbean corals: Roles of turbidity, temperature, and nutrients. Palaios 1994, 9, 576–598. [Google Scholar] [CrossRef]
  80. Webster, J.M.; Braga, J.C.; Humblet, M.; Potts, D.C.; Iryu, Y.; Yokoyama, Y.; Fujita, K.; Bourillot, R.; Esat, T.M.; Fallon, S.; et al. Response of the Great Barrier Reef to sea-level and environmental changes over the past 30,000 years. Nat. Geosci. 2018, 11, 426–432. [Google Scholar] [CrossRef] [Green Version]
  81. Kiessling, W.; Simpson, C.; Beck, B.; Mewis, H.; Pandolfi, J.M. Equatorial decline of reef corals during the last Pleistocene interglacial. Proc. Natl. Acad. Sci. USA 2012, 109, 21378–21383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Cheng, H.; Edwards, R.L.; Sinha, A.; Spötl, C.; Yi, L.; Chen, S.; Kelly, M.; Kathayat, G.; Wang, X.; Li, X.; et al. The Asian monsoon over the past 640,000 years and ice age terminations. Nature 2016, 534, 640–646. [Google Scholar] [CrossRef] [PubMed]
  83. Gradstein, F.M.; Ogg, J.G.; Schmitz, M.D.; Ogg, G.M. Geologic Time Scale 2020; Elsevier: Amsterdam, The Netherlands, 2020; p. 1357. [Google Scholar] [CrossRef]
  84. Naish, T.; Powell, R.; Levy, R.; Wilson, G.; Scherer, R.; Talarico, F.; Krissek, L.; Niessen, F.; Pompilio, M.; Wilson, T. Obliquity-paced Pliocene West Antarctic ice sheet oscillations. Nature 2009, 458, 322–328. [Google Scholar] [CrossRef] [PubMed]
  85. Huang, H.; Gao, Y.; Ma, C.; Jones, M.M.; Zeeden, C.; Ibarra, D.E.; 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, eabf9489. [Google Scholar] [CrossRef]
Figure 1. Study area and the location of core XK-1. (a) The South China Sea (SCS). (b) Topography of Xisha (Paracel) Islands. (c) The skeleton of the Shidao Island with the five formations identified in regional studies. The age boundary of each formation was from [18], and the skeleton (c) was modified from previous reports [26,34,35]. The base map data were generated using the open and free software DIVA–GIS 7.5 (http://www.diva-gis.org/, accessed on 10 September 2021).
Figure 1. Study area and the location of core XK-1. (a) The South China Sea (SCS). (b) Topography of Xisha (Paracel) Islands. (c) The skeleton of the Shidao Island with the five formations identified in regional studies. The age boundary of each formation was from [18], and the skeleton (c) was modified from previous reports [26,34,35]. The base map data were generated using the open and free software DIVA–GIS 7.5 (http://www.diva-gis.org/, accessed on 10 September 2021).
Jmse 09 01031 g001
Figure 2. Profile of the studied section of core XK-1. (a) Lithological changes. (b) Sampling position. (c) The preliminary age-depth model of core XK-1. (d) Comparison between orbital eccentricity on 405 kyr band and magnetic susceptibility (MS) of core XK-1 on the preliminary timescale. The data presented in (c,d) were from Reference [18].
Figure 2. Profile of the studied section of core XK-1. (a) Lithological changes. (b) Sampling position. (c) The preliminary age-depth model of core XK-1. (d) Comparison between orbital eccentricity on 405 kyr band and magnetic susceptibility (MS) of core XK-1 on the preliminary timescale. The data presented in (c,d) were from Reference [18].
Jmse 09 01031 g002
Figure 3. Changes in the magnetic parameters of core XK-1.
Figure 3. Changes in the magnetic parameters of core XK-1.
Jmse 09 01031 g003
Figure 4. Relationship between the magnetic parameters of core XK-1.
Figure 4. Relationship between the magnetic parameters of core XK-1.
Jmse 09 01031 g004
Figure 5. Magnetic properties of biogenic reefs in core XK-1. (a) Hysteresis loops at 1256 m. (b) Hysteresis loops at 1239 m. (c) FORC diagram at 1239 m with a smoothing factor of 9. (d) Scanning result of the sample at 1239 m with a 200-μm interval. Note that hysteresis loop measurements were taken of 11 samples in total, and, since the other 9 samples had a similar loop with (a), these loops are not shown in the figure.
Figure 5. Magnetic properties of biogenic reefs in core XK-1. (a) Hysteresis loops at 1256 m. (b) Hysteresis loops at 1239 m. (c) FORC diagram at 1239 m with a smoothing factor of 9. (d) Scanning result of the sample at 1239 m with a 200-μm interval. Note that hysteresis loop measurements were taken of 11 samples in total, and, since the other 9 samples had a similar loop with (a), these loops are not shown in the figure.
Jmse 09 01031 g005
Figure 6. Magnetic proxy (MagXK) of biogenic reefs in core XK-1. (a) MagXK (red bold line) with magnetic parameters (dashed thin lines), versus eccentricity on a 405 kyr band (dashed bold line) [55]. (b) The evolution spectrum of the MagXK. (c) The squared wavelet coherence [56] between MagXK and eccentricity. The thick black contour designates the 5% significance level against red noise, and the cone of influence (COI) where edge effects might distort the picture is shown as a lighter shade. Arrows: in-phase pointing right, anti-phase pointing left.
Figure 6. Magnetic proxy (MagXK) of biogenic reefs in core XK-1. (a) MagXK (red bold line) with magnetic parameters (dashed thin lines), versus eccentricity on a 405 kyr band (dashed bold line) [55]. (b) The evolution spectrum of the MagXK. (c) The squared wavelet coherence [56] between MagXK and eccentricity. The thick black contour designates the 5% significance level against red noise, and the cone of influence (COI) where edge effects might distort the picture is shown as a lighter shade. Arrows: in-phase pointing right, anti-phase pointing left.
Jmse 09 01031 g006
Figure 7. Comparison between the MagXK of core XK-1 and various paleoenvironmental proxies. (af) The MagXK of core XK-1 versus global δ18O and δ13C records [16]. (c,g) The MagXK versus the CLP-ZL MS, the loess magnetic susceptibility of Zhuanglang Profile in the Chinese Loess Plateau [61]. (d,h) The MagXK versus global sea level [62]. Dashed lines, original data; bold lines on the left panel, low-passed-filtered component (FFT filter, >0.5 Myr); lines on the right panel, high-passed-filtered component (FFT filter, <0.5 Myr).
Figure 7. Comparison between the MagXK of core XK-1 and various paleoenvironmental proxies. (af) The MagXK of core XK-1 versus global δ18O and δ13C records [16]. (c,g) The MagXK versus the CLP-ZL MS, the loess magnetic susceptibility of Zhuanglang Profile in the Chinese Loess Plateau [61]. (d,h) The MagXK versus global sea level [62]. Dashed lines, original data; bold lines on the left panel, low-passed-filtered component (FFT filter, >0.5 Myr); lines on the right panel, high-passed-filtered component (FFT filter, <0.5 Myr).
Jmse 09 01031 g007
Figure 8. Comparison between paleoenvironmental proxies. All data were from low-passed-filtered components (FFT filter >1.0 Myr). See Figure 6 and Figure 7.
Figure 8. Comparison between paleoenvironmental proxies. All data were from low-passed-filtered components (FFT filter >1.0 Myr). See Figure 6 and Figure 7.
Jmse 09 01031 g008
Table 1. Results of a principal component analysis of magnetic proxies.
Table 1. Results of a principal component analysis of magnetic proxies.
Component 1Extraction Sums of Squared Loadings
Total% of VarianceCumulative %
MagXK3.5187.7783.77
Not included0.256.2794.03
0.225.3999.42
0.020.58100
1 Four concentration-dependent magnetic parameters (χARM and IRMs) were analyzed and are shown in Figure 6.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, Y.; Liu, X.; Chen, W.; Yi, L. Magnetic Properties and Initiation of Biogenic Reefs in Xisha Islands, South China Sea, at the Oligo–Miocene Boundary. J. Mar. Sci. Eng. 2021, 9, 1031. https://doi.org/10.3390/jmse9091031

AMA Style

Li Y, Liu X, Chen W, Yi L. Magnetic Properties and Initiation of Biogenic Reefs in Xisha Islands, South China Sea, at the Oligo–Miocene Boundary. Journal of Marine Science and Engineering. 2021; 9(9):1031. https://doi.org/10.3390/jmse9091031

Chicago/Turabian Style

Li, Yibing, Xinyu Liu, Weiwei Chen, and Liang Yi. 2021. "Magnetic Properties and Initiation of Biogenic Reefs in Xisha Islands, South China Sea, at the Oligo–Miocene Boundary" Journal of Marine Science and Engineering 9, no. 9: 1031. https://doi.org/10.3390/jmse9091031

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