Next Article in Journal
Research on Maneuverability Prediction of Double Waterjet Propulsion High Speed Planing Craft
Previous Article in Journal
Characterization of Intertidal Macrofaunal Communities of Two Sandy Beaches under Different Anthropogenic Pressures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Response of Magnetic Minerals to the Mid-Brunhes Climate Event Recorded in Deep-Sea Sediments of the West Philippine Sea

1
Key Laboratory of Virtual Geographic Environment, School of Geography, Nanjing Normal University, Ministry of Education, Nanjing 210023, China
2
State Key Laboratory of Marine Geology, Tongji University, Shanghai 200092, China
3
Key Laboratory of Muddy Coastal Geo-Environment, Tianjin Centre, China Geological Survey, Tianjin 300170, China
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2022, 10(12), 1977; https://doi.org/10.3390/jmse10121977
Submission received: 15 November 2022 / Revised: 4 December 2022 / Accepted: 6 December 2022 / Published: 12 December 2022
(This article belongs to the Section Geological Oceanography)

Abstract

:
The Mid-Brunhes Event (MBE) was one of the most important global climate events since 800 ka. The deep-sea palaeoceanographic changes in the Western Pacific might have been more sensitive to the MBE and they have not been well documented yet. In this study, we investigated a deep-sea core collected from the Western Philippine Sea and then obtained the magnetism record since around 900 ka. The MBE signal in the deposition process was derived from the concentration-dependent (χ, χARM and SIRM) and the grain-size-dependent magnetic parameters (χARM/χ and χARM/SIRM). Across the MBE, the content of magnetic minerals and the fluctuation of the grain size decreased. Both the content and grain-size indices calculated from magnetic parameters are well correlated with the glacial–interglacial alternations and display a major shift at the MBE. The decreased grain size of magnetic minerals may be associated with the flourishing biota in the tropical Pacific during interglacial intervals. The accumulation of magnetic minerals in the Western Philippine Sea was mainly regulated by the sediment accumulation rate, which may be related to the shoaling of the carbonate compensation depth. Overall, this study expanded the environmental magnetism record for the MBE, suggested a new possible influence of this critical climatic event on the deep-sea deposition process in Western Philippine Sea and inferred the interactions between various environmental systems on glacial–interglacial timescales.

1. Introduction

The Philippine Sea, located in the Western Pacific, is a key area for material and energy exchanges. It plays an important role in regional source-to-sink processes and global climate changes [1,2,3]. Reconstructing the paleoenvironment in the Philippine Sea on geological timescales (since around 900 ka) can provide us with useful insights for revealing the deep-sea process in the Western Pacific and its linkage to global climate changes.
The Mid-Brunhes Event (MBE) was one of the most important global climate events since 800 ka [4]. It was characterized by remarkably increased amplitudes of frequent glacial–interglacial cycles and carbonate changes since ~430 ka [5,6]. According to the ice core records, the Antarctic temperature was cooler and the concentration of atmospheric CO2 was lower during the interglacial intervals before the MBE [7,8]. It was also revealed by deep-sea benthic δ18O that the sea level was higher in the interglacial period after the MBE [9]. The MBE started with warm and high productivity events, and it was most prominent in the southern Arctic region. However, it seems that the concentration of atmospheric CH4 and the Asian monsoon were less affected by the MBE [10,11,12]. Thus, the exploration of the mechanism of the MBE is becoming an increasingly significant issue [4,13,14,15].
There are many controversies about the corresponding mechanism of the MBE [4,13,14,15]. It was pointed out that the meteorite impact event that occurred at 480 ka in Antarctica may have led to the occurrence of the MBE [16]. Some researchers believe that the MBE may have something to do with the latitudinal shift of the Southern Hemisphere westerlies which led to the CO2 respiring from the Southern Ocean [17] and/or a slowdown of Antarctic Bottom Water (AABW) formation 14. However, the relationships between orbital forcing, greenhouse gases and ice volume are complex [9]. In addition, it is not well known how the global carbon cycle changed before and after the MBE. Thus, more studies need to be carried out to better understand the mechanism of the MBE and the earth’s climate response to the MBE.
In this study, we investigated a 3.5 m deep-sea sediment (Core I8) collected from the Western Philippine Sea, Western Pacific (Figure 1). The chronology of the section has been well set up by Xu et al. (2022) [18]. Reconstructing the accumulate history of the magnetic minerals in Core I8 can potentially unlock a wealth of new information from the deep-sea sediment. The MBE signal in the deposition process was derived from the magnetism record of Core I8 since around 900 ka. The purpose of this study is twofold: (1) to deduce the deep-sea palaeoceanographic processes according to the changes of magnetic properties; and (2) to discuss the linkage of the deep-sea process in the Western Philippine Sea to the global climate changes.

2. Materials and Methods

2.1. The Study Core (I8)

The study core, I8 (12°75′ N, 131°29′ E, water depth 5405 m), was collected from the western part of the Philippine Sea (Figure 1) using a gravity corer. The core was drilled to a depth of 3.5 m. The sediments of core I8 are generally light brown to brown in color and mainly contain silty clay. For paleomagnetic and paleoenvironment studies, we obtained 172 samples in 2 cm intervals by non-magnetic plastic cubes (2 × 2 × 2 cm3). Prior to this, Xu et al. (2022) [18] determined two magnetic zones in core I8 (Figure 2a,b) and established a reliable chronostratigraphic framework by magnetostratigraphy and tuning the Ba intensities of the I8 core to the global ice volume (Figure 2d,e). The sediment accumulation rates (SARs) for each section were then calculated (Figure 2f). Average SAR was estimated as 0.41 ± 0.15 cm/ka over the depositional interval. As seen in Figure 2f, the SAR has shown a general downward trend since ~887 ka with a great jump at about 450 ka and it has remained at a low level after 400 ka.

2.2. Magnetic Measurements

Magnetic susceptibility (χlf and χhf) was measured using an AGICO MFK1-FA multi-frequency Kappabridge magnetic susceptibility meter at frequencies of 976 Hz and 15,616 Hz, respectively. By using a 2G enterprises SQUID magnetometer with inline AF coils, anhysteretic remanent magnetization (ARM) was imparted to all the samples with a peak alternating field (AF) of 100 mT and a direct biasing field of 0.05 mT. Isothermal remanent magnetization (IRM) was induced with a 2-G Enterprises model 660 pulse magnetizer successively in pulsed fields of 1 T, which is regarded here as saturated isothermal remanent magnetization (SIRM). Then, the IRM of 0.3 T was applied in the opposite direction, denoted IRM−0.3T. The percentage of frequency-dependent susceptibility (χfd%), hard isothermal remanent magnetization (HIRM) and S-ratio were calculated using the following equations [21]. All the magnetic measurements were performed in the State Key Laboratory of Marine Geology, Tongji University.
χ fd = χ lf χ hf χ lf × 100 % ,
HIRM = SIRM IRM 0.3 T 2 ,
S 300 = S I R M I R M 0.3 T 2 × S I R M × 100 % ,
S Ratio = IRM 0.3 T SIRM .

3. Results and Analyses

3.1. Changes in Magnetic Parameters

The low-frequency magnetic susceptibility (χlf) of core I8 varies from 9.32 × 10−7 m3/kg to 1.52 × 10−6 m3/kg with an average value of 1.21 × 10−6 m3/kg and a standard deviation of 1.63 × 10−7 m3/kg. The high-frequency magnetic susceptibility (χhf) varies from 8.89 × 10−7 m3/kg to 1.45 × 10−6 m3/kg with an average value of 1.15 × 10−6 m3/kg and a standard deviation of 1.56 × 10−7 m3/kg (Table 1). The susceptibility of ARM (χARM) is usually sensitive to single-domain (SD) grains [22,23], and SIRM can be used to infer magnetic particles excluding the influence of superparamagnetic (SP) grains [24]. The values of χARM and SIRM are 3.85 × 10−6−6.91 × 10−6 m3/kg and 1.57 × 10−2−2.68 × 10−2 Am2/kg, respectively (Table 1). Their average values are 5.25 × 10−6 m3/kg and 2.05 × 10−2 Am2/kg and both of them display a similar pattern to χ curves (Figure 3a–d). The χ, χARM and SIRM generally reflect the concentration of magnetic minerals, especially ferrimagnetic minerals [25]. The concentration-dependent magnetic parameters (χ, χARM and SIRM) have similar variations throughout the core, which suggests that these proxies were not greatly influenced by mineralogical variations. They all declined greatly at 480 ka, indicating a significant decrease in magnetic minerals.
Overall, the χfd% has a relatively low value, indicating a small amount of super-paramagnetic material. The minimum and maximum values of χfd% are 3.44% and 4.81%, and the average value is about 4.26%. The standard deviation of χfd% is 4.26% (Table 1). Unlike the χlf and the χhf, the χfd% does not show an obvious downward trend at around 480 ka. However, the variation of χfd% has a greater amplitude after ~480 ka (Figure 3e). The χARM/χ varies between 3.70 and 4.37 with an average value of 4.37, and χARM/SIRM varies between 2.11 × 10−4 m/A and 2.56 × 10−4 m/A with an average value of 4.37 m/A. Their standard deviations are 0.31 and 2.05 × 10−4 m/A, respectively (Table 1). Both the χARM/χ and χARM/SIRM are commonly used as grain size indicators for ferrimagnetic minerals, as they decrease with increasing grain size [25]. The main discrepancy is that the first proxy is more sensitive to size change but may be biased by magnetic susceptibility from super-paramagnetic and paramagnetic material. The latter is less sensitive than the first, but it only contains remanence, avoiding the effect of super-paramagnetic and paramagnetic grains. In addition, the χARM/χ and χARM/SIRM agree well with less deflection by super-paramagnetic and paramagnetic grains, both of them are main proxies of magnetic grain size (Figure 3f,g).
The S-ratio reflects the relative proportion of low- to high-coercivity minerals in samples. The larger the ratio is, the more low-coercivity magnetic minerals there are, and this kind of mineralogy change can indicate source variation. The S-ratio of Core I8 varies between 0.9–1 with an average value of 0.97 and the standard deviation is 0.02, indicating the dominance of low-coercivity magnetic minerals throughout the whole sequence (Figure 3h).
Principal component analysis was used to obtain the more representative concentration- and grain-size-dependent magnetic parameters. The results show that χ and χARM can better reflect the magnetic mineral content and that the χARM/χ and χARM/SIRM are more relevant to the magnetic mineral grain size (Table 1). Based on the principal component analysis, we detrended, normalized and then averaged the χlf and SIRM data to build the magnetic mineral content index, and perform the same operation on χARM/χ and χARM/SIRM to obtain the magnetic mineral grain-size index.

3.2. Relationships between Magnetic Parameters

Based on the general variation in the concentration-dependent (χ, χARM and SIRM) and the grain-size-dependent magnetic parameters (χARM/χ and χARM/SIRM), the magnetic series can be divided into two stages (Figure 3). In Stage 1 (before 480 ka), the magnetic mineral content was at a high level and accompanied by large vibrations. Meanwhile, the grain-size-dependent magnetic parameters have a higher frequency of variation during this time. Stage 2 (since 480 ka) was characterized by a sudden drop of magnetic mineral content and it has remained at a low level since then. At the same time, there was a slowdown in the frequency of changes in grain size. The difference between the two stages may be related to the MBE.
In order to determine the variation of magnetic parameters and its linkage to the MBE, we analyzed the correlation coefficients of different parameters for both stages (before and after 480 ka). As shown in their scatter plots (Figure 4), the differences associated with both the concentration- and grain-size-dependent parameters are significant before and after MBE. Specifically, in the case of concentration-dependent parameters (χlf vs. SIRM), the differences in the slopes and in regression intercepts are both evident (Figure 4a). As seen in Figure 4b, it is the same case with the grain-size-dependent parameters (χARM/χ vs. χARM/SIRM). For the case of χARM vs. χARM/χ, the differences in regression intercepts are evident with a small difference in the slopes (Figure 4c). Both the slopes and regression intercepts are significantly different in the case of SIRM vs. χARM/SIRM (Figure 4d). Additionally, for other paired parameters, such as χARM vs. SIRM and χfd% vs. χARM/SIRM, they are not shown in this work because of the poor fitting. The distinct regression intercepts indicate that the basic background of magnetic particles would be evident, while a difference in regression slopes may indicate different magnetic properties in different time intervals. Hence, it is inferred that magnetic properties changed significantly at ~480 ka and the MBE was a key factor for deep-sea environmental changes in the West Philippine Sea.

3.3. Power Spectral Analyses

In order to obtain the environmental significance hidden in these magnetic parameters, we performed spectral analysis on the proxies in this study and other widely used environmental records. The power density spectra of Benthic δ18O stack (LR04) [26] was calculated by Acycle software (https://acycle.org/, accessed on 1 November 2022) in MATLAB [27,28], as were the mean grain size (MGS) of the Chinese Loess Plateau [10], the magnetic parameters (ꭓlf,ARM, ꭓARM/ꭓ, ꭓARM/SIRM), and both the content and grain-size indices of magnetic minerals.
LR04, one of the most widely used proxies for global climate changes, was first calculated for its power spectrum. There appeared all the well-known peaks corresponding to eccentricity (100 ka), obliquity (40 ka) and precession (23 ka) when considering the interval since 900 ka (Figure 5a). The eccentricity cycle shows the glacial–interglacial variation related to the deep-water temperature and sea level change [29]. In the studies of tropical oceans, the obliquity and precession cycle are usually used as indicators of monsoonal changing [30,31]. MGS has been widely used to indicate the intensity of the Asian monsoon, and it shows significant peaks at both 41 ka and 23 ka (Figure 5a).
The magnetic mineral parameters related to the content (ꭓ, SIRM; Figure 5b) show obvious periods of 100 ka, 41 ka and 23 ka in the spectrum. Of course, such a phenomenon also appears in the more representative magnetic mineral content index (Figure 5d). However, the grain-size-dependent magnetic parameters (ꭓARM/ꭓ, ꭓARM/SIRM; Figure 5c) and the calculated magnetic mineral grain-size index only show a period of 100 ka in the spectrum.

4. Discussion

It has been reported that the so-called MBE, representing an enlargement of glacial–interglacial alternations since ~430 ka, has had a major influence on key climatic processes. To better understand the potential linkages between regional paleoenvironmental processes and global climate changes, the magnetic properties of core I8 were further compared with various proxies (Figure 6). All these proxies exhibited in Figure 6 have been broadly employed for paleoenvironmental inferences in different types of sediments [32,33,34,35,36].
As discussed in the Section 3.3, both the content and grain-size indices of magnetic minerals have evident cycles (Figure 5 and Figure 6a,b). Moreover, their changing patterns are similar to those of the Northern Hemisphere glaciation, as reflected in the stacked benthic δ18O record of deep-sea sediments [26] (LR04, Figure 6b). The similarity between the magnetic mineral index of core I8 and the LR04 record, indicating that they were both regulated by eccentricity period (~100 ka, Figure 5), indicates the force of glacial–interglacial periods. We know that the grain-size-dependent magnetic parameters (ꭓARM/ꭓ, ꭓARM/SIRM; Figure 5c) and the calculated grain-size index only show a period of 100 ka in the spectrum. However, the content index also exhibits a more evident cycle in precession bands (~23 ka, Figure 6a). It generally agrees with changes in the Asian monsoon as reflected in the stack grain size (MGS) of the China Loess Plateau [10] (Figure 6c), especially during interglacial intervals. Spectral analysis confirms this consistency between them (Figure 5), thus highlighting an integrated forcing of Asian monsoon variability on regional sedimentary processes, perhaps related to provenance.
The magnetic mineral content index was usually higher during the interglacial periods, which indicates a higher amount of magnetic minerals in the sediments (Figure 6a,b). Meanwhile, the higher magnetic mineral grain-size index during the interglacial periods represents the thinning of magnetic minerals. In general, during the interglacial periods, the magnetic minerals in the sediments of core I8 mainly show an increase in content and decrease in grain size. Nevertheless, for the entire deposition phase, the indices of magnetic minerals underwent a significant change in the MBE. After this crucial global climate event, the change frequency and amplitude of the magnetic mineral grain-size index gradually slows down and the magnetic minerals become finer. Due to the climate becoming warmer in the interglacial period after the MBE, magnetic mineral grain-size thinning may be associated with flourishing biota in the tropical Pacific during interglacial intervals [37,38,39,40,41].
The SAR of Core I8 dropped greatly after the MBE (Figure 2f), which is consistent with the content-dependent magnetic parameters (Figure 6a–d). The low SAR during the post-MBE period may be the key reason for the reduction of magnetic mineral content. The lower sedimentation rate led to the large reduction of magnetic minerals after the MBE. As one of the most important global climate events, the MBE had a great influence on the carbonate compensation depth (CCD) [42,43]. By comparing the paleoenvironmental proxies, Xu et al. (2022) [18] proposed that the CCD was about 200 m shallower than in the post-MBE time. The CCD shoaling was reflected in the bulk carbon δ13C of core C6 [18] (Figure 6e) and the weaker AABW [44,45] recorded at ODP 1088 [46] (Figure 6f), inferring a weakening of deep-sea effective sedimentation leading to the decrease in SAR and magnetic mineral content.

5. Conclusions

Based on a well-dated record of deep-sea sediment in the Western Philippine Sea, environmental magnetic information for the middle Pleistocene was derived to deduce the deep-sea palaeoceanographic processes. Our main findings are summarized as follows: (1) The magnetic series can be divided into two stages. Before the MBE, the magnetic mineral content was at a high level and accompanied by large vibrations. Meanwhile, the grain-size-dependent magnetic parameters have a higher frequency of variation during this time. After the MBE, there was a sudden drop in magnetic mineral content and it has remained at a low level since then. Furthermore, there was a slowdown in the frequency of changes in grain size. (2) Both content and grain-size indices calculated by the magnetic parameters can be well correlated to the glacial–interglacial alternations and display a major shift at the MBE. (3) The decrease in magnetic mineral grain size may be associated with flourishing biota in the tropical Pacific during interglacial intervals. (4) The accumulation of magnetic minerals in the Western Philippine Sea was mainly regulated by the sediment accumulation rate, which may be related to the carbonate compensation depth. According to these observations, this study expanded the environmental magnetism record of the MBE, suggested a new possible influence of the MBE on the deposition process in the Western Philippine Sea and inferred interactions between various environmental systems on glacial–interglacial timescales.

Author Contributions

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

Funding

This research was funded by the National Programme on Global Change and Air-Sea Interaction, grant number GASI-04-HYDZ-02; and the National Natural Science Foundation of China, grant number 42177422.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Huang, J.; Wan, S.; Zhang, G.; Xu, Z.; Liu, H.; Dong, J.; Li, A.; Li, T. Impact of seafloor topography on distribution of clay minerals in the east Philippines Sea. Mar. Geol. Quat. Geol. 2017, 37, 77–85. (In Chinese) [Google Scholar] [CrossRef]
  2. Yu, Z.; Wan, S.; Colin, C.; Yan, H.; Bonneau, L.; Liu, Z.; Song, L.; Sun, H.; Xu, Z.; Jiang, X.; et al. Co-evolution of monsoonalprecipitation in east Asia and the tropical Pacific ENSO system since2.36 Ma: New insights from high-resolution clay mineral records in thewest Philippine Sea. Earth Planet. Sci. Lett. 2016, 446, 45–55. [Google Scholar] [CrossRef]
  3. Jiang, F.; Frank, M.; Li, T.; Chen, T.; Xu, Z.; Li, A. Asian dust input in the western Philippine Sea: Evidence from radiogenic Sr and Nd isotopes. Geochem. Geophys. Geosyst. 2013, 14, 1538–1551. [Google Scholar] [CrossRef] [Green Version]
  4. Jansen, J.; Kuijpers, A.; Troelstra, S. A Mid-Brunhes climatic event: Long-term changes in global atmosphere and ocean circulation. Science 1986, 232, 619–622. [Google Scholar] [CrossRef] [PubMed]
  5. Crowley, T. Late Quaternary carbonate changes in the North Atlantic and Atlantic/Pacific comparisons. In The Carbon Cycle and Atmospheric CO2: Natural Variations Archean to Present; Geophysical Monograph Series; Sundquist, E.T., Broecker, W.S., Eds.; AGU: Washington, DC, USA, 1985; Volume 32, pp. 271–284. [Google Scholar]
  6. Wang, P.; Li, Q.; Tian, J.; Jian, Z.; Liu, C.; Li, L.; Ma, W. Long-term cycles in the carbon reservoir of the Quaternary ocean: A perspective from the South China Sea. Natl. Sci. Rev. 2014, 1, 119–143. [Google Scholar] [CrossRef] [Green Version]
  7. Jouzel, J.; Massondelmotte, V.; Cattani, O.; Dreyfus, G.; Falourd, S.; Hoffmann, G.; Minster, B.; Nouet, J.; Barnola, J.; Chappellaz, J. Orbital and Millennial Antarctic Climate Variability over the Past 800,000 Years. Science 2007, 317, 793–796. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Lüthi, D.; Floch, M.; Bereiter, B.; Blunier, T.; Barnola, J.-M.; Siegenthaler, U.; Raynaud, D.; Jouzel, J.; Fischer, H.; Kawamura, K.; et al. High-resolution carbon dioxide concentration record 650,000–800,000 years before present. Nature 2008, 453, 379–382. [Google Scholar] [CrossRef] [Green Version]
  9. Barth, A.; Clark, P.; Bill, N.; He, F.; Pisias, N. Climate evolution across the Mid-Brunhes Transition. Clim. Past 2018, 14, 2071–2087. [Google Scholar] [CrossRef] [Green Version]
  10. Sun, Y.; Clemens, S.; 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]
  11. Loulergue, L.; Schilt, A.; Spahni, R.; Massondelmotte, V.; Blunier, T.; Lemieux, B.; Barnola, J.; Raynaud, D.; Stocker, T.; Chappellaz, J. Orbital and millennial-scale features of atmospheric CH4 over the past 800,000 years. Nature 2008, 453, 383–386. [Google Scholar] [CrossRef]
  12. Cheng, H.; Edwards, R.; 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]
  13. Paillard, D. The timing of Pleistocene glaciations from a simple multiple-state climate model. Nature 1998, 391, 378–381. [Google Scholar] [CrossRef]
  14. Yin, Q. Insolation-induced mid-Brunhes transition in Southern Ocean ventilation and deep-ocean temperature. Nature 2013, 494, 222–225. [Google Scholar] [CrossRef]
  15. Tzedakis, P.; Crucifix, M.; Mitsui, T.; Wolff, E. A simple rule to determine which insolation cycles lead to interglacials. Nature 2017, 542, 427–432. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Rice, A.; Weihaupt, J.; Vand, H. Multiple meteoroid impacts in Antarctica at 481,000ky: A possible cause for the mid-Brunhes Event/MIS 11 Stage via the disruption of the West Antarctic Ice Sheet; Agu Fall Meeting. In AGU Fall Meeting Abstracts; NH51C-1249; AGU: Washington, DC, USA, 2010. [Google Scholar]
  17. Kemp, A.; Grigorov, I.; Pearce, R.; Naveira Garabato, A. Migration of the Antarctic Polar Front through the mid-Pleistocene transition: Evidence and climatic implications. Quat. Sci. Rev. 2010, 29, 1993–2009. [Google Scholar] [CrossRef] [Green Version]
  18. Xu, D.; Yi, L.; Yuan, H.; Chen, W. Geochronological Evidence Inferring Carbonate Compensation Depth Shoaling in the Philippine Sea after the Mid-Brunhes Event. J. Mar. Sci. Eng. 2022, 10, 745. [Google Scholar] [CrossRef]
  19. Kim, S.; Wessel, P. Directional median filtering for regional-residual separation of bathymetry. Geochem. Geophys. Geosyst. 2008, 9, Q03005. [Google Scholar] [CrossRef] [Green Version]
  20. Wessel, P.; Smith, W.H.F.; Scharroo, R.; Luis, J.; Wobbe, F. Generic Mapping Tools: Improved Version Released. EOS Trans. Am. Geophys. Union 2013, 94, 409–410. [Google Scholar] [CrossRef] [Green Version]
  21. Tauxe, L.; Butler, R.; Van Der Voo, R.; Banerjee, S. Essentials of Paleomagnetism; University of California Press: Berkeley, CA, USA, 2010; pp. 1–512. [Google Scholar] [CrossRef]
  22. Maher, B. Magnetic properties of some synthetic sub-micron magnetites. Geophys. J. 1988, 94, 83–96. [Google Scholar] [CrossRef]
  23. Duan, Z.; Gao, X.; Liu, Q. Anhysteretic remanent magnetization (ARM) and its application to geoscience. Prog. Geophys. 2012, 27, 1929–1938. (In Chinese) [Google Scholar] [CrossRef]
  24. Evans, M.; Heller, F. Environmental Magnetism: Principles and Applications of Enviromagnetics; Academic Press: Amsterdam, The Netherlands, 2003; pp. 1–299. [Google Scholar] [CrossRef]
  25. Thompson, R.; Oldfield, F. Environmental Magnetism; George Alien & Unwin: London, UK, 1986; pp. 1–166. [Google Scholar] [CrossRef]
  26. Lisiecki, L.; Raymo, M. A Pliocene-Pleistocene stack of 57 globally distributed benthic δ18O records. Paleoceanography 2005, 20, PA1003. [Google Scholar] [CrossRef] [Green Version]
  27. Li, M.; Hinnov, L.; Kump, L. Acycle: Time-series analysis software for paleoclimate research and education. Comput. Geosci. 2019, 127, 12–22. [Google Scholar] [CrossRef]
  28. Li, M.; Huang, C.; Hinnov, L.; Ogg, J.; Chen, Z.-Q.; Zhang, Y. Obliquity-forced climate during the Early Triassic hothouse in China. Geology 2016, 44, 623–626. [Google Scholar] [CrossRef]
  29. Lisiecki, L.; Raymo, M. Plio–Pleistocene climate evolution: Trends and transitions in glacial cycle dynamics. Quat. Sci. Rev. 2007, 26, 56–69. [Google Scholar] [CrossRef] [Green Version]
  30. Prell, W.; Kutzbach, J. Monsoon variability over the past 150,000 years. J. Geophys. Res. 1987, 92, 8411–8425. [Google Scholar] [CrossRef]
  31. Lee, S.; Poulsen, C. Tropical Pacific climate response to obliquity forcing in the Pleistocene. Paleoceanography 2005, 20, PA4010. [Google Scholar] [CrossRef]
  32. Huang, W.; Jian, Z.; Bühring, C. The Millennial-scale climate fluctuations revealed by the records of color reflectance from ODP site 1144 in the northern south China sea. Mar. Geol. Quat. Geol. 2003, 3, 5–10. [Google Scholar] [CrossRef]
  33. Li, Y.; Chen, Y.; Jiang, X.; Yi, L. Color Reflectance of Coastal Sediments in the South Bohai Sea and its Implication to Orbital Forcing of East Asian Summer Monsoon. Front. Earth Sci. 2021, 9, 944. [Google Scholar] [CrossRef]
  34. Liu, L.; Zheng, H.; Jian, Z. Visible Reflectance Record of South China Sea Sediments during the Past 220ka and Its Implications for East Asian Monsoon Variation. Earth Sci. 2005, 5, 543–549. [Google Scholar] [CrossRef]
  35. Peng, S.; Guo, Z. Brightness characteristics of the Late Tertiary laterite record of the Western Peaks. Quat. Sci. 2003, 1, 110. [Google Scholar]
  36. Yi, L.; Chen, S.; Ortiz, J.; Chen, G.; Peng, J.; Liu, F.; Chen, Y.; Deng, C. 1500-year cycle dominated Holocene dynamics of the Yellow River delta, China. Holocene 2015, 26, 222–234. [Google Scholar] [CrossRef] [Green Version]
  37. Yamazaki, T. Environmental magnetism of Pleistocene sediments in the North Pacific and Ontong-Java Plateau: Temporal variations of detrital and biogenic components. Geochem. Geophys. Geosystems 2009, 10, Q07Z04. [Google Scholar] [CrossRef]
  38. Liu, S.; Deng, C.; Xiao, J.; Li, J.; Paterson, G.; Chang, L.; Yi, L.; Qin, H.; Pan, Y.; Zhu, R. Insolation driven biomagnetic response to the Holocene warm period in semi-arid East Asia. Sci. Rep. 2015, 5, 8001. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Chang, L.; Harrison, R.; Zeng, F.; Berndt, T.; Roberts, A.; Heslop, D.; Zhao, X. Coupled microbial bloom and oxygenation decline recorded by magnetofossils during the Palaeocene–Eocene Thermal Maximum. Nat. Commun. 2018, 9, 4007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. 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, 1–11. [Google Scholar] [CrossRef]
  41. Yi, L.; Wang, H.; Deng, X.; Yuan, H.; Xu, D.; Yao, H. Geochronology and Geochemical Properties of Mid-Pleistocene Sediments on the Caiwei Guyot in the Northwest Pacific Imply a Surface-to-Deep Linkage. J. Mar. Sci. Eng. 2021, 9, 253. [Google Scholar] [CrossRef]
  42. Woosley, R. Carbonate Compensation Depth. In Encyclopedia of Geochemistry: A Comprehensive Reference Source on the Chemistry of the Earth; White, W.M., Ed.; Springer International Publishing: Cham, Switzerland, 2018; pp. 204–205. [Google Scholar]
  43. Pälike, H.; Lyle, M.; Nishi, H.; Raffi, I.; Ridgwell, A.; Gamage, K.; Klaus, A.; Acton, G.; Anderson, L.; Backman, J.; et al. A Cenozoic record of the equatorial Pacific carbonate compensation depth. Nature 2012, 488, 609–614. [Google Scholar] [CrossRef] [Green Version]
  44. Clark, P.; Alley, R.; Pollard, D. Northern Hemisphere Ice-Sheet Influences on Global Climate Change. Science 1999, 286, 1104–1111. [Google Scholar] [CrossRef]
  45. Brook, E.; Buizert, C. Antarctic and global climate history viewed from ice cores. Nature 2018, 558, 200–208. [Google Scholar] [CrossRef]
  46. Hodell, D.; Venz-Curtis, K. Late Neogene history of deepwater ventilation in the Southern Ocean. Geochem. Geophys. Geosyst. 2006, 7, Q09001. [Google Scholar] [CrossRef]
Figure 1. Schematic map showing the core location (red star) and oceanographic setting. AABW: Antarctic Bottom Water, LCDW: Lower Circumpolar Deep Water. The map was generated by using Generic Mapping Tools (GMT toolbox, http://gmt.soest.hawaii.edu/, accessed on 1 November 2022) in C programs [19,20].
Figure 1. Schematic map showing the core location (red star) and oceanographic setting. AABW: Antarctic Bottom Water, LCDW: Lower Circumpolar Deep Water. The map was generated by using Generic Mapping Tools (GMT toolbox, http://gmt.soest.hawaii.edu/, accessed on 1 November 2022) in C programs [19,20].
Jmse 10 01977 g001
Figure 2. Chronostratigraphic framework of core I8 [18]. (a), ChRM inclination of core I8; (b), polarity of core I8; (c), geological polarity timescale (GPT2020); (d,e) tuning Ba intensities to LR04 δ18O record, M/B boundary is the boundary of the magnetic Matuyama and Brunes chronons (~773 ka), green numbers 1–25 represent the marine isotope stages; (f), sediment accumulation rate (SAR).
Figure 2. Chronostratigraphic framework of core I8 [18]. (a), ChRM inclination of core I8; (b), polarity of core I8; (c), geological polarity timescale (GPT2020); (d,e) tuning Ba intensities to LR04 δ18O record, M/B boundary is the boundary of the magnetic Matuyama and Brunes chronons (~773 ka), green numbers 1–25 represent the marine isotope stages; (f), sediment accumulation rate (SAR).
Jmse 10 01977 g002
Figure 3. Changes in magnetic parameters of Core I8 since around 900 ka; (a), the low-frequency magnetic susceptibility (χlf); (b), the high-frequency magnetic susceptibility (χhf); (c), the susceptibility of anhysteretic remanent magnetization (χARM); (d), the saturated isothermal remanent magnetization (SIRM); (e), the percentage of frequency-dependent susceptibility (χfd%); the grain-size-dependent magnetic parameters χARM/χ (f) and χARM/SIRM (g); (h), the S-ratio. It can be divided into two stages (before and after 480 ka) according to the changes in magnetic parameters.
Figure 3. Changes in magnetic parameters of Core I8 since around 900 ka; (a), the low-frequency magnetic susceptibility (χlf); (b), the high-frequency magnetic susceptibility (χhf); (c), the susceptibility of anhysteretic remanent magnetization (χARM); (d), the saturated isothermal remanent magnetization (SIRM); (e), the percentage of frequency-dependent susceptibility (χfd%); the grain-size-dependent magnetic parameters χARM/χ (f) and χARM/SIRM (g); (h), the S-ratio. It can be divided into two stages (before and after 480 ka) according to the changes in magnetic parameters.
Jmse 10 01977 g003
Figure 4. Relationships between magnetic parameters of Core I8. (a), χlf vs. SIRM; (b), χARM/χ vs. χARM/SIRM; (c), χARM vs. χARM/χ; (d), SIRM vs. χARM/SIRM. The scatter plots of magnetic parameters before (pink circle) and after 480 ka (blue triangle) are presented. Next to them are the regression equation and correlation coefficient of two stages.
Figure 4. Relationships between magnetic parameters of Core I8. (a), χlf vs. SIRM; (b), χARM/χ vs. χARM/SIRM; (c), χARM vs. χARM/χ; (d), SIRM vs. χARM/SIRM. The scatter plots of magnetic parameters before (pink circle) and after 480 ka (blue triangle) are presented. Next to them are the regression equation and correlation coefficient of two stages.
Jmse 10 01977 g004
Figure 5. Power spectral calculation by Acycle software [27,28]. (a), Benthic δ18O stack LR04 [26] and mean grain size of Chinese Loess Plateau [10]; (b), ꭓlf and ꭓARM; (c), ꭓARM/ꭓ and ꭓARM/SIRM; (d) magnetic mineral content index and magnetic mineral grain-size index. CL: confidence level.
Figure 5. Power spectral calculation by Acycle software [27,28]. (a), Benthic δ18O stack LR04 [26] and mean grain size of Chinese Loess Plateau [10]; (b), ꭓlf and ꭓARM; (c), ꭓARM/ꭓ and ꭓARM/SIRM; (d) magnetic mineral content index and magnetic mineral grain-size index. CL: confidence level.
Jmse 10 01977 g005
Figure 6. Comparison of various environmental proxies; (a), magnetic content index (calculated by ꭓlf and SIRM); (b), the benthic δ18O stack LR04 [26]; (c), stack grain size (MGS) of China Loess Plateau (CLP) [10]; (d), magnetic minerals grain-size index (calculated by χARM/χ and χARM/SIRM); (e), the bulk carbon δ13C of core C6 in Western Philippine Sea [18], indicated CCD shoaling; (f), Benthic δ13C records of ODP Sites 1088, indicating changes in Circumpolar Deep Water [46]. MIS, marine isotope stages, which are labelled as numbers 5–21 on the top.
Figure 6. Comparison of various environmental proxies; (a), magnetic content index (calculated by ꭓlf and SIRM); (b), the benthic δ18O stack LR04 [26]; (c), stack grain size (MGS) of China Loess Plateau (CLP) [10]; (d), magnetic minerals grain-size index (calculated by χARM/χ and χARM/SIRM); (e), the bulk carbon δ13C of core C6 in Western Philippine Sea [18], indicated CCD shoaling; (f), Benthic δ13C records of ODP Sites 1088, indicating changes in Circumpolar Deep Water [46]. MIS, marine isotope stages, which are labelled as numbers 5–21 on the top.
Jmse 10 01977 g006
Table 1. Results of magnetic parameters.
Table 1. Results of magnetic parameters.
MinMaxAverageStandard DeviationPrincipal Component Analysis
ContentGrain Size
lf (m3/kg)9.32 × 10−71.52 × 10−61.21 × 10−61.63 × 10−70.973
hf (m3/kg)8.89 × 10−71.45 × 10−61.15 × 10−61.56 × 10−70.972
SIRM(Am2/kg)1.57 × 10−22.68 × 10−22.05 × 10−22.73 × 10−30.648
χARM (m3/kg)3.85 × 10−66.91 × 10−65.25 × 10−67.07 × 10−70.908
fd (%)3.444.814.260.240.455
χARM3.704.894.370.310.945
χARM /SIRM(m/A)2.11 × 10−43.03 × 10−42.56 × 10−42.05 × 10−40.937
S-Ratio0.901.000.970.02
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cai, Y.; Li, Y.; Jiang, X. Response of Magnetic Minerals to the Mid-Brunhes Climate Event Recorded in Deep-Sea Sediments of the West Philippine Sea. J. Mar. Sci. Eng. 2022, 10, 1977. https://doi.org/10.3390/jmse10121977

AMA Style

Cai Y, Li Y, Jiang X. Response of Magnetic Minerals to the Mid-Brunhes Climate Event Recorded in Deep-Sea Sediments of the West Philippine Sea. Journal of Marine Science and Engineering. 2022; 10(12):1977. https://doi.org/10.3390/jmse10121977

Chicago/Turabian Style

Cai, Yun, Yibing Li, and Xingyu Jiang. 2022. "Response of Magnetic Minerals to the Mid-Brunhes Climate Event Recorded in Deep-Sea Sediments of the West Philippine Sea" Journal of Marine Science and Engineering 10, no. 12: 1977. https://doi.org/10.3390/jmse10121977

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